steady-state modeling of extrusion cast film process, neck

126
1 STEADY-STATE MODELING OF EXTRUSION CAST FILM PROCESS, NECK-IN PHENOMENON AND RELATED EXPERIMENTAL RESEARCH: A REVIEW Tomas Barborik, Martin Zatloukal* Polymer Centre, Faculty of Technology, Tomas Bata University in Zlin, Vavreckova 275, 760 01 Zlin, Czech Republic Keywords: Modeling of polymer processing, polymers, rheology and fluid dynamics, polymer flows, flat film production, neck-in phenomenon. *Corresponding author: [email protected] This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset. PLEASE CITE THIS ARTICLE AS DOI: 10.1063/5.0004589

Upload: others

Post on 30-Jan-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

1

STEADY-STATE MODELING OF EXTRUSION CAST FILM

PROCESS, NECK-IN PHENOMENON AND RELATED

EXPERIMENTAL RESEARCH: A REVIEW

Tomas Barborik, Martin Zatloukal*

Polymer Centre, Faculty of Technology, Tomas Bata University in Zlin,

Vavreckova 275, 760 01 Zlin, Czech Republic

Keywords: Modeling of polymer processing, polymers, rheology and fluid dynamics,

polymer flows, flat film production, neck-in phenomenon.

*Corresponding author: [email protected]

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

2

ABSTRACT

This review paper provides the current state of knowledge of steady-state modeling of the

extrusion cast film process used to produce flat polymer films, as well as related experimental

research with a particular focus on the flow instability neck-in. All kinematic models used (i.e.

1, 1.5, 2 and 3-dimensional models) together with the utilized constitutive equations, boundary

conditions, simplified assumptions and numerical methods are carefully summarized. The

effect of draw ratio, Deborah number (i.e. melt relaxation time related to experimental time),

film cooling, second to first normal stress difference ratio at the die exit, uniaxial extensional

strain hardening and planar-to-uniaxial extensional viscosity ratio on the neck-in is discussed.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

3

1 INTRODUCTION

Extrusion film casting is an industrially important process, which in practice has a solid

place among polymer processing technologies. It can be classified as a continuous, high-speed

manufacturing process, during which monolayer or co-extruded multilayer thin, highly oriented

films are produced. A wide range of the plastic films and sheets produced by this technology

are used in many different applications of daily and technical use: plastic bags, consumer

packaging, magnetic tapes for storing audio video content, optical membranes for liquid crystal

displays, flexible electronics, foils for capacitors and microporous membranes used primarily

in separation processes (from microfiltration to reverse osmosis or as separators in lithium-ion

batteries for mobile devices and electric vehicles [1–4]) or as a product for further processing

by other technologies such as thermoforming and biaxial orientation [5, 6].

The growing demand for the quantity production and quality of manufactured films,

together with the introduction of new materials, requires new approaches in production line. Of

particular interest is to reach desirable properties of the produced films and to keep film

thickness uniform and width as close as possible to the designed extrusion die width. In order

to eliminate an expensive and time-consuming trial-and-error approach widely used in the

plastics industry to optimize film casting process, one can use a computer modeling for the

optimization of die design and process conditions for a given polymer system. This strategy can

provide a better insight into the problem, broaden the knowledge on relationships between

process/rheological variables and propose possible approaches to deal with them to optimize

the process or provide a better understanding of basic underlying mechanics [6].

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

4

1.1 Film casting process description

The extrusion film casting is a technology, in which polymer pellets are conveyed,

homogenized, compressed and melted in an extruder. Then, the polymer melt is pushed through

the uniform slit die (center-fed T die or coat-hanger die) with typically about 1–2 mm gap size

[5]. The thick sheet is then intensively stretched in the machine direction using a constant rotary

whose circumferential velocity, vx(X), is higher than the average polymer melt velocity at the

die exit, vx(0). This leads to the orientation of macromolecules and reduced film thickness, and

due to a sufficiently high cooling rate, the final film dimensions are fixed. Intensity of the

stretching is given by a draw ratio, which is defined as ( ) ( )x xDR v X v 0= . Additionally, an

increase in DR, cooling rate or stretching distance can cause temperature and/or stress induced

crystallization, which can enhance final film properties. The process is visualized in Figure 1.

At the chill roll, several other technological devices can be used to provide a better contact

line between the film and the chill roll and to increase the heat transfer rate, such as an air knife

(a slit nozzle blows a jet of cooled air to film) or electrostatic pinning [5–7]. In the latter device,

a high voltage wire is positioned parallel to the grounded chill roll that generates an electrostatic

discharge exerting electrostatic force on the film to increase the film-chill roll contact. Another

alternative with the similar result is a vacuum box, which provides a vacuum between the film

and the chill roll [7–9]. In addition to cooling on the chill roll, the polymer film is naturally

cooled to some extent, depending on the length of the drawing zone, by passing through the

surrounding environment. This can be enhanced by introducing convection air or an inert gas

source into this section or by passing the film through a fluid bath [10]. Additionally, the

produced polymeric film can also be subjected to treatment (plasma treating, heating and biaxial

orientation) depending on the desired properties and purpose of the final product. Polymer

behavior and extensional conditions in the drawing zone have been shown to be key factors

determining the final mechanical and optical properties of the film [6, 11].

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

5

To produce highly functional films with tailored properties, multiple layers of different

polymer melts can be coextruded and stretched, i.e., the properties of the film are given by each

individual layer. In this way, multilayer films with enhanced properties, such as oxygen and

moisture impermeability, strength, chemical resistance or color can be produced [12]. An

alternative continuous film production technology is called the extrusion film blowing process.

In this process, the extruded tube is inflated by the internal pressure into a bubble shape having

a thin wall thickness, which is simultaneously quenched and hauled off [13–17]. In contrast to

this competing film production technology, films made by extrusion film casting have good

transparency, uniformity of thickness, a smoother surface and are produced at a higher

production rate [6].

According to the current industry practice, where a wide variety of films are produced with

a requirement for use in heterogeneous applications, manufactures process a broad range of

materials by using film casting technology. Frequently used polymeric materials include

low-density polyethylene, LDPE; high-density polyethylene, HDPE; linear low-density

polyethylene, LLDPE; polypropylene, PP; polyethylene terephthalate, PET; and polystyrene,

PS. The extrusion film casting is also suitable for low viscosity polymers [18] and

biodegradable polymers such as for example polylactide (PLA) or its blends with polybutylene

succinate (PBS) or poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PBAT) [19, 20]. Since

these films have a wide range of applications, there is a requirement to produce a wide range of

sizes. The film width can typically range from 0.1 m to 10 m, thicknesses from 20 μm to

2000 μm [11] at production rates ranging from 70 to 200 m/min. Tolerable thickness variation

is reported to be from 3 to 5 % [5]. The plastics industry, which focuses on the production of

plastic foils, is currently undergoing a major change, due to the gradual transition from

conventional commodity polymers to more advanced [6]. These include, for example,

metallocene polymers with an easily modifiable structure which make it possible to

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

6

significantly improve the final properties of the film. Structural polymers such as polyethylene

terephthalate, polycarbonate, polyamide, polyphenylsulfide have become popular materials for

producing films with high heat resistance. The line speed for the production of polymer films

is gradually increasing for economic reasons and in some cases (e.g. polypropylene or

polyethylene terephthalate) may reach up to 500 m/min [6].

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

7

2 FLOW INSTABILITIES

The presence of an air-polymer interface in the drawing zone makes it possible to develop

various types of flow instabilities that severely limit the desired film quality and quantity. Their

formation is influenced by processing conditions, heat transfer and rheology of the processed

polymer. For example, if the draw ratio reaches some critical value (for the given process

conditions, die design and polymer used), transient hydrodynamic instability, called draw

resonance, begins to occur [21]. This instability causes oscillations of the film dimensions,

although the volumetric flow supplied from the slit die and take-up speed are kept constant, see

Figure 2. These periodic fluctuations in film width and thickness (measured in the center of the

film) are offset by half-wavelength (i.e. maximum in width corresponds to the minimum

thickness) and vice versa [22]. Extension of the drawing distance, increased cooling effects and

the use of polymers with strong extensional strain hardening can stabilize the process and move

the onset of draw resonance toward higher draw ratios [23].

Film breakage is another feature that can be observed during increasing the draw ratio. In

this case, the chains cannot be reorganized to relieve local stresses within the time frame

imposed by the deformation, resulting in a cohesive failure between the polymer chains and

disintegration of the film. This can be seen in polymers containing long chain branches or a

high molecular weight portion processed at high line speeds and cooling rates, leading to good

process stability but also to the development of high tensile stress [6].

Neck-in and edge-beading are flow phenomena which are the most common instabilities in

the production of flat films because they occur and destabilize the flow at any processing

conditions. These instabilities are described and reviewed below in greater detail.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

8

2.1 Neck-in

Upon leaving the die, the extruded polymer in the form of a thick sheet exhibits swelling

due to its viscoelastic nature. This relaxation of molecular stress is then influenced by the

velocity rearrangement that occurs during the transition from a confined shear flow in the slit

die to the downstream extension. When the polymer sheet is hauled off further downstream and

stable processing conditions are met, its cross-sectional dimensions are monotonically reduced

due to the external drawing force exerted on the sheet by the rotary winding drum. In addition

to the desirable reduction in film thickness, the width of the film is reduced. This defect is called

the neck-in and can be defined as the difference between the half-width of the film at the die

exit and the final half-width of the solidified film (Figure 3). The neck-in is considered to be a

typical instability occurring in extensional flows as explained by Larson in [24], even if it occurs

under steady state extrusion conditions (i.e. that the stress and the velocity are not time

dependent at the given point of the stretched film) because it can have serious consequences

since it might lead to breakage of the film.

The role of extrudate swell on the film drawing was investigated by using viscoelastic

constitutive equations such as the Leonov model [25] and the linear PTT model [26]. It was

demonstrated that the negative value of the second normal stress difference causes swelling in

the thickness direction much higher than in the width direction of the extrudate [26]. Even if

the intensity of the extrudate swell rapidly decreases by increased take-up velocity, there might

be “a certain amount of swelling persisting near the die exit” lowering the melt velocity at this

region [26]. This can increase the melt orientation because the actual DR “expressed in terms

of the velocity at the point of the film’s maximum thickness” is higher than the conventional

DR based on the melt velocity at the die exit [26].

Based on experimental studies (Table 1) and theoretical analyses (Tables 2–5), the

following material parameters and process variables have been identified to have a significant

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

9

impact on the neck-in phenomenon: molecular weight (Mw), molecular weight distribution

(MWD), relaxation time (), the ratio of the second and first normal stress difference at the die

exit (−N2/N1), long chain branching (LCB), strain hardening in uniaxial extension (SH in E,U),

planar to uniaxial extensional viscosity ratio (E,P/E,U), draw ratio, DR, take-up length, X, take-

up rate, vx(X), melt speed at the die exit, vx(0), and temperature T. The role of each individual

parameter on this phenomenon is summarized in Table 6. In order to clarify the reading of Table

6, let us provide here an example explaining its first line, which should be read as follows: In

1986, Dobroth and Erwin reported that the neck-in for LDPE increases if the draw ratio, DR,

(adjusted via the average polymer melt velocity at the die exit, vx(X)) increases or if the take-

up length, X, increases.

As can be seen, the reduction of the neck-in can be achieved by an increasing the polymer

melt relaxation time (via broadening MWD and/or increasing Mw and/or decreasing T),

increasing the melt speed at the die exit vx(0) (maintaining a constant DR) or reducing the

air-gap (distance between die and roll), X. All these three variables determine the elasticity of

the melt, which can be evaluated in terms of the Deborah number defined as

( )xv 0

DeX

= (1)

It is obvious, that if the Deborah number (i.e. melt elasticity) increases, the neck-in decreases

(although the stability of the process in terms of the maximum attainable draw ratio, DR, at

which the film breaks may be lowered [27–29]). Thus, it is appropriate to maintain the level of

elasticity reasonably high to minimize neck-in, which can be achieved by increasing the

relaxation time and/or the melt speed at the die exit or by reducing the air-gap (see Eq. 1). The

effect of relaxation time on the neck-in phenomenon determined experimentally for two linear

low-density polyethylenes, LLDPEs [30], and two linear polypropylenes, PPs [31–34], is

provided in Figures 4–5. It is important to mention that a different definition of relaxation time

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

10

can be found in the reviewed literature. In the studies based on single-mode constitutive

equations, the utilized Maxwell relaxation time [27], the shortest [35] or characteristic

relaxation time (determined by the reciprocal frequency at the intersection of the storage

modulus G’ and the loss modulus G’’ curves [36] or by fitting the strain rate dependent steady

uniaxial extensional viscosity data [37, 38]) are typically used to calculate De. In the case of

multi-mode constitutive equations, the relaxation time for each mode [39–44] or an average

relaxation time, , is calculated to determine De by using the following expression [45–47]:

N2

j j

j 1

N

j j

i 1

G

G

=

=

=

(2)

where j and Gj is the relaxation time and the modulus, respectively, in the jth relaxation mode.

Also, in some experimental studies, the longest relaxation time ( 0

0 EJ = , where 0 is the zero-

shear viscosity and 0

EJ is the linear steady-state elastic compliance) [33] or the characteristic

(reptation-mode) relaxation time representing the onset of shear-thinning [30], are used.

The role of DR on the neck-in is complex, depending whether the polymeric chains are

linear or branched or if DR is changed via vx(X) or vx(0). Current experimental studies showed

that for linear polymers (i.e. for PP, PET, LLDPE, HDPE), an increase in DR (by an increase

in vx(X)) always increases the neck-in but for branched polymers (such as LDPEs), interestingly,

the trend can even become an opposite (see Table 6 and Figure 6). This unexpected trend was

attributed to the strain-hardening behavior of LDPE in elongational flow [48]. The situation

becomes also complex, if DR is increased by reduction in vx(0) (keeping the vx(X) constant). It

was reported for branched LDPE and linear PET that an increase in DR (by decrease in vx(0))

reduces the neck-in, but for linear isotactic polypropylene (iPP) the trend was found to be

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

11

surprisingly opposite for given processing conditions. This unexpected trend for iPP was

attributed to the increased crystallization rate, which caused quicker film solidification.

It was found that the introduction of strain hardening, SH, in uniaxial extensional viscosity,

E,U, by incorporating long chain branches into polymer backbone chains, reduces the neck-in

phenomenon. This trend is illustrated in Figure 7 for linear low-density polyethylene and highly

branched low-density polyethylene. Seay and Baird [30] revealed that the addition of sparse

long chain branching (LCB) to polymer chains, i.e. SH in E,U, is more significant for film width

conservation than broadening the molecular weight distribution (MWD). In addition, they have

found that increasing LCB of long and short chains reduced the neck-in at low and high draw

ratios, respectively. This effect is sometimes used to improve final width for films made from

polymers prone to the neck-in (such as HDPE, LLDPE) using coextrusion technology in which

the surface/edge portion of the film is made of a material having a long-chain branching (such

as LDPE) and a core from linear polymer [49–51].

Recent viscoelastic modeling of the extrusion film casting process, which followed the

corresponding neck-in measurements, suggests that reduction of E,P/E,U or −N2/N1 at the die

exit (if De > 0.1) can also reduce the neck-in phenomenon (see Table 6).

In order to understand the role of extensional rheology and the die exit stress state, it is

necessary to discuss the mechanism and physical background of the neck-in in more detail. Ito

et al. [52], performed an experimental study on metallocene-catalyzed linear low-density

polyethylene, mLLDPE, aimed to visualize the flow in the air-gap region during a film casting

operation. They used small aluminum particles and made streamline measurement by using the

particle tracking method. These particles were placed across the film width at the die exit (one

particle for each measurement) and their movement was monitored by a camera for different

draw ratios (DR = 4.4, 7.5 and 12.2). It has been found that streamlines at the film center are

straight regardless the drawing intensity (i.e. there is planar extensional flow) whereas the

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

12

streamlines at the near-edge region were found to be curved with a tapered transversal spacing

in the flow direction confirming presence of uniaxial extensional flow (see Figure 8). Moreover,

an increase in the draw ratio caused an increase in the streamlines tapering, which lead to a

more pronounced neck-in phenomenon. Therefore, if E,U increases due to SH in such a way

that E,P/E,U is decreased (i.e. if the resistance against the uniaxial extensional flow becomes

much higher than the resistance against the planar extensional flow), the polymer melt starts to

prefer the planar extensional flow at the expense of the uniaxial elongation flow and thus, the

neck-in is decreased. Similarly, the reduction in −N2/N1 at the die exit physically means an

induction of a planar prestretch inside the extrusion die (for example by using converging

instead of a parallel flow channel), which (if remembered by the melts, i.e. if De > 0.1) increases

dominance of the planar extensional flow in the post die area and therefore, the neck-in is

reduced.

In the industrial practice, it is useful to have a tool that can provide a reasonable evaluation

of the neck-in for a particular polymeric material and processing conditions prior to film

production itself, where its determination via a trial-and-error approach can be very expensive.

It is therefore not surprising that considerable efforts have been made to relate the neck-in with

the air-gap, extensional strain rate and relaxation time [53] (Eqs. 3–4), both planar and uniaxial

extensional viscosities [54] (Eq. 5), and to the strain hardening, the ratio of planar to uniaxial

extensional viscosity, the Deborah number and the die exit stress state [38] (Eq. 6). These

simple analytical models are easy to use and have the advantage to gain a correlation between

the particular model variables with the naked eye in order to identify key process and material

parameters to optimize them for neck-in reduction. In more detail, Ito et al. [53] in 2003

developed a model based on the Dobroth-Erwin model [55], which assumes a planar extensional

flow in the middle of the film and a uniaxial extensional flow at the edge. According to their

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

13

theory, the final film width is determined by the ratio of planar viscosities in the axial and

transverse directions with respect to the flow. Proposed relationship for the neck-in, NI,

considering that the polymer melt behaves as a Newtonian fluid, is following

1

NI X2

= (3)

where the air-gap, X, is the only variable (i.e. draw ratio, deformation rate, relaxation time or

viscosity are not included). The model was tested using experimental data for linear HDPE,

short chain branched LLDPE and long chain branched LDPE melts expressed as the neck-in

plotted against the air-gap at four different draw ratios. The model was shown to correctly

predict the general trend between NI and X, i.e. an increase in X causes an increase in NI, and

the predicted slope of the theoretical line was close to experimental data for HDPE and LLDPE

at short air-gap values and the highest draw ratios at given processing conditions. On the other

hand, the model tended to overpredict NI (especially for LDPE) without the ability to predict

the effect of draw ratio on NI, as expected due to the absence of deformation rate and any

rheological parameters in Eq. 3. If the upper convected Maxwell model is used, the expression

for NI yields the following form

( )p

1NI 1 2 X

2= − (4)

where p denotes the extensional strain rate of the planar part defined as ( )=p pd t / dt

( p is the Hencky strain of the planar deformation) and λ is the characteristic relaxation time.

Although the neck-in trend predictions are consistent with the observations with respect to the

melt elasticity or the air-gap, the model unrealistically predicts neck-in decrease for the

increased draw-down ratio, which was attributed to the used constitutive equation. Further

works have moved further and tried to predict neck-in based on uniaxial and planar extensional

viscosities. Shiromoto et al. [54] in 2010 developed a theoretical model based on the force

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

14

balance and film deformation in the post die area. The authors found that NI can be correlated

with the planar to uniaxial extensional viscosity ratio rather than with the strain hardening in

uniaxial extension or with the ratio of planar viscosities. These findings were transformed into

the following expression for NI

0.5

E,P

E,U

NI X

(5)

where ηE,P and ηE,U denotes planar and uniaxial extensional viscosity. Eq. 5 was validated using

relevant experimental data for three long chain branched LDPEs having different Mw and

MWD. The ratio ηE,P/ηE,U was determined for all three samples using the multi-mode

exponential type of Phan–Thien and Tanner constitutive model utilizing parameters identified

on the experimental data obtained from rotational and capillary rheometers. ηE,U at low

deformation rates and high extensional rates was determined with a Meissner-type rheometer

(ARES-EVF, TA Instruments) and the Cogswell method [56], respectively. The key limitation

of the Eq. 5 is the absence of variables allowing to evaluate the role of the Deborah number and

the die exit stress state on NI. Barborik and Zatloukal [37, 38] continued the research of the

neck-in phenomenon in the period 2017-2018 with respect to the ratio of the second and first

normal stress difference at the die exit, −N2/N1, uniaxial extensional strain hardening,

ηE,U,max/3η0, melt elasticity (captured via the Deborah number, De) and the ratio of planar-to-

uniaxial extensional viscosity, ηE,P/ηE,U. Using an isothermal 1.5-dimensional (1.5D) membrane

model and viscoelastic modified Leonov constitutive equation, the following expression for

maximum attainable neck-in was proposed:

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

15

( )

( )

( )

2.113

7.43

0.514

0.593 1 exp 1073.742 1

1 1.027arctan 0.849X

NI

arctan 0.514 tanh 3.953

1.027 1 exp 0.849

− − − +

+

= +

− −

E,P

E,U

2

1

E,U,maxE,U,max

00

ηDe

η

N

N

ηηDe

3η3η

De

(6)

The predictions of the model were found to be in good agreement with the corresponding

experimental data [38] (see Figure 9) capturing all the trends obtained numerically, i.e. that NI

is reduced if

• −N2/N1 at the die exit decreases (i.e. for increased planar pre-stretch of the melt inside

the extrusion die),

• De increases,

• ηE,P/ηE,U decreases,

• ηE,U,max/3η0 increases.

It has also been revealed that there is a threshold of about 0.1 for De above which the neck-in

phenomenon starts to be strongly dependent on the −N2/N1 ratio at the die exit. In other words,

if De > 0.1, the flow history inside the die (i.e. the die design) starts to significantly affect the

neck-in phenomenon. It is important to mention that Eq. 6 represents an analytical

approximation of numerical solutions based on an isothermal (1.5D) membrane model utilizing

the modified Leonov constitutive equation (single mode) for the processing conditions, in

which the maximum attainable neck-in is achieved (i.e. for very high draw ratios only) were

0.011 De 0.253 , E,P

E,U

η0.825 1.910

η ,

E,U,max

0

η2.047 10.096

3η and

2

1

N0.017 0.680

N − . The basic form of Eq. 6 has been derived from the assumed linear

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

16

function between NI and E,P E ,U/ , in which its constants were allowed to vary with De

according to the Avrami exponential functions. Detailed derivation of Eq. 6 is provided in [37,

38]. Validation of Eq. 6 was performed for different highly branched LDPEs. All rheological

quantities appearing in Eq. 6 were predicted by the single-mode modified Leonov model whose

parameters were identified from uniaxial extensional viscosity data only. This procedure seems

to be reasonable at least for the given LDPEs and applied processing conditions. In order to

experimentally evaluate ηE,P/ηE,U, one could use the Cogswell model and measured entrance

pressure drops on capillary rheometer by using circular and rectangle capillaries [57, 58]

whereas −N2/N1 can be evaluated using Han’s methods utilizing exit pressure drop measures

by using a capillary rheometer equipped by a slit die [59–64]. The key advantage of Eq. 6 in

contrary to Eqs. 3–5 is a consideration of uniaxial and planar extensional viscositites along with

the Deborah number and die exit stress state (quantified via −N2/N1). On the other hand, the

model is only applicable to very high draw ratios and does not take into account the full

relaxation spectrum, film cooling and crystallization, which can be considered as its key

limitations.

2.2 Edge-beading

In addition to the neck-in phenomenon, an interrelated defect, referred to as the edge-beading

or the dog-bone defect, is formed making the edge portions of the film substantially thicker

than its central part (Figure 10). The size of these raised parts can be five times higher compared

to the center and several centimeters wide. The predominant cause of the edge-beads formation

is the edge-stress effect [55] arising due to the neck-in phenomenon and it’s intensity increases

with the draw ratio (see Figure 11). The following equation was derived in [55] to evaluate the

edge-beading:

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

17

edge

f

center

f

hB DR

h= = (7)

where B is the bead ratio and edge

fh and center

fh represents edge and center final thickness,

respectively. This equation was derived by simply comparing the strains of the center

(undergoing planar stretch) and edge (undergoing uniaxial stretch) elements between the roll

and the die neglecting surface tension, extrudate swell and assuming melt incompressibility (i.e.

without the need to use any constitutive equation). Equation 7 was successfully validated for

LDPE for DR between 1 and 20 [55] (see Figure 12). It has also been shown (when comparing

simulations based on the Newtonian and UCM models) that increasing the melt elasticity (by

increasing the Deborah number) decreases the intensity of the edge-beading [28].

The raised edges are often trimmed with a slit razer, scrapped and optionally reprocessed

to achieve a uniform film surface. Regardless of the large amount of waste material, the

occurrence of edge-beads also causes air entrapment between the film and the chill roll, which

in turn results in poorer film quality. In the manufacturing practice, a technological procedure

of opening lateral parts of the extrusion slit die (i.e. the gap size at the edge is bigger than in the

center) can be found in order to create thicker edges that would restrain the neck-in level in

comparison to the situations when the edge-beads and the neck-in would evolved in the natural

way [11].

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

18

3 MATHEMATICAL MODELING OF THE EXTRUSION FILM CASTING

PROCESS

The drawing of polymer films or filaments has taken an enormous amount of attention and

has been extensively studied both experimentally and theoretically in the past four decades

because of its great importance in the polymer processing industry.

3.1 Flow kinematics

Individual research groups focused on experimental works [30, 48, 52, 55, 65–74] dealing

with flow visualization, effects of temperature, crystallization, molecular weight distribution or

long chain branching on kinetics of the film casting process (see Table 1). Theoretical research

has not disappeared, and attention has been drawn to the development and use of numerical

models (primarily considering steady-state conditions) of different dimensionality such as

1D [22, 25, 26, 45, 75–85], 1.5D [18, 27, 47, 53, 86–93, 29, 94–100, 35, 37, 38, 41–44], 2D [8,

12, 28, 36, 39, 40, 46, 54, 101–113] and full 3D [114–116] (see Tables 2–5) using different

types of constitutive equations taking into account non-isothermal effects, crystallization,

inertia, and gravity. The 1D model here is based on the assumption of an infinite film width and

assumes the following velocity field

( )x xv v x= (8)

yv 0= (9)

xz

vv z

x

= −

(10)

that is, the flow deformation in the drawing region is planar [52, 55]. The 1.5D model is simply

the 1D model with variable film width proposed in [27, 28]. This simplified model, which

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

19

retains the ability to cover the reduction in film width in the drawing length while reducing the

dimensionality of the solved problem, is based on the assumption that all velocity components

are an exclusive function of the drawing length position, x, and vary linearly with respect to its

corresponding direction. In this case, a velocity field is assumed in form of

( )x xv v x= (11)

( )yv yf x= (12)

( )zv zg x= − (13)

The 2D approximation developed in [112, 113] is based on the so-called membrane

hypothesis considering that one dimension of the film is small compared to the others [112].

The film thickness is much smaller (several orders of magnitude) than the film width and the

take-up length, so it can be assumed that the velocity component in the machine and the

transversal direction are independent of thickness direction, i.e., uniform across the thickness.

The reduced velocity field is given in the following form

( )x xv v x, y= (14)

( )y yv v x, y= (15)

yx

z

vvv z

x y

= − +

(16)

The 3D model utilizes velocity components without any restriction, which are given bellow:

( )x xv v x, y, z= (17)

( )y yv v x, y, z= (18)

( )z zv v x, y, z= (19)

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

20

3.2 Constitutive equations

Different types of constitutive equations were used to model film casting, as shown in

Tables 2–5. They are introduced and briefly discussed in this chapter. For simplicity, they are

usually provided in a single mode version. Note that in the multi-mode approach, a

discontinuous relaxation spectrum is used and the stress tensor is expressed as N

jj 1=

= where

j represents the stress tensor in the jth mode.

3.2.1 Newtonian model

This constitutive equation describes the behavior of ideal Newtonian fluids by Eq. 20.

02 D = (20)

Here is the extra stress tensor, 0 is the Newtonian shear viscosity (zero-shear-rate viscosity)

and D is the deformation rate tensor defined as

( )T

v v v

1D L L , L v

2= + = (21)

where v represents the velocity field, T denotes the transpose of the tensor and is the gradient

operator. The Newtonian model predicts constant steady shear viscosity (0), steady uniaxial

(30) and planar (40) extensional viscosities, which is correct also for polymer melts, but only

at very low extensional rates, where they behave as Newtonian fluids. The key advantage of

this model is its mathematical simplicity and utilization of only one adjustable parameter, 0,

which can be determined from simple shear flow measurements. On the other hand, the model

does not have the ability to describe the elasticity and memory of fluids, the extensional strain

thinning or the extensional strain hardening, typically occurring for polymer melts at medium

and high extensional strain rates. It has been found that the model provides reasonable NI values

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

21

only at low DR values, it predicts essentially parabolic thickness profile across the film width

as well as increased NI values with increased DR [105]. The convergence is almost guaranteed

[89]. At high DRs, the Newtonian model predicts artificially high neck-in, and there are also

discrepancies between experiments and temperature profile predictions, as shown for PET in

[90] (see Figure 13). This constitutive equation was used in the following studies: 1.D (4 works)

[80], [82], [22], [26]; 1.5D (10 works) [18], [100], [86], [87], [88], [27], [89], [90], [53], [97];

2D (10 works) [112], [12], [28], [101], [102], [39], [104], [105], [113], [111]; 3D (1 work)

[114].

3.2.2 Generalized Newtonian model

The generalized Newtonian model is simply the Newtonian model given by Eq. 20, in which

0 is replaced by a viscosity scalar function ( )DII , which is allowed to vary with the second

invariant of deformation rate tensor DII defined as ( )22tr D . In the simple shear flow, 2

DII =

, uniaxial extensional flow, 2

DII 3= , and planar extensional flow, 2

DII 4= . Here and

represents shear and extensional strain rate, respectively.

Power-law (or Ostwald–de Waele) model [117–119]

( ) ( )n 1

D DII m II−

= (22)

where m (the flow consistency index) and n (the flow behavior index, which is lower than 1 for

polymer melts) are adjustable parameters. This model allows to model shear, uniaxial and

planar extensional viscosities plotted as the function of deformation rates as a simple line in

log-log scale with the slope equal to n-1. The key advantage of this model its mathematical

simplicity and low number of used parameters. A key disadvantage of this model is over

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

22

prediction of shear and extensional viscosities at low deformation rates, the incapability to

describe a smooth transition from Newtonian to non-Newtonian flow regime and the

incapability to represent fluid elasticity, memory and extensional rheology. It was found that

model predictions start to significantly deviate from the Newtonian solution when the power-

law index n becomes less than 0.8 [113]. This model was utilized in the following studies within

this review: 1.D (1 work) [81]; 1.5D (0); 2D (2 works) [113], [8]; 3D (0).

Cross model [120]

( )( )

0D a

D

II

1 II

− = +

+

(23)

The model was used in: 1.D (0); 1.5D (3 works) [91], [93], [96]; 2D (1 work) [105]; 3D (0).

Carreau model [121]

( )

( )

0D 1 n

2 2

D

II

1 II

− = +

+

(24)

The model was used in: 2D (2 works) [110], [111]; 3D (1 work) [115].

Carreau-Yasuda model [122]

( )

( )

0D 1 n

a a

D

II

1 II

− = +

+

(25)

The model was used in: 1.D (1 work) [79]; 1.5D (0) ; 2D (1 work) [40]; 3D (0).

The above viscosity models use the following parameters: η0 (zero-shear-rate viscosity), η∞

(infinite-shear-rate viscosity), λ (relaxation time), a (characterizes the sharpness of the

transition from Newtonian to non-Newtonian flow regime) and n (characterizes the slope

between viscosity and deformation rate in a log-log scale) are model parameters. Utilization of

low number of parameters, mathematical simplicity, capability to represent steady shear

viscosity of polymer melts in a wide range of shear rates and correct predictions of steady

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

23

extensional viscosities at low extensional rates (i.e. equal to 30 and 40 for uniaxial and planar

extensional viscosities, respectively) represent the advantages of these models. The main

disadvantages are the inability to represent fluid memory and extensional rheology for branched

polymers.

The use of generalized Newtonian models in film casting modeling has made it possible to

capture some very important trends observed experimentally, namely that NI intensity and the

edge bead increase with DR or that an increase in planar to uniaxial extensional viscosity ratio

increases NI, in agreement with the viscoelastic PTT model [40], [110].

3.2.3 Upper-Convected Maxwell (UCM) model

One of the simplest model allowing to represent some basic viscoelastic features of polymer

melts is called the Upper-Convected Maxwell model, which is given by the Eq. 26.

02 D

+ = (26)

As can be seen, the key difference between the Newtonian and the Upper-Convected Maxwell

model is the elastic term

, which consists of the relaxation time, , and the upper-convected

time derivative of the stress tensor,

, defined as

( )T

v vv L L

t

= + − −

(27)

The key advantage of this model is its mathematical simplicity, low number of used parameters

(, 0, or alternatively and the elastic modulus G where 0 = G), it takes into account the

melt memory and the first normal stress difference, N1, is predicted to be nonzero.

Disadvantages: The model predicts unrealistically strong extensional strain hardening without

the ability to predict the extensional strain thinning and it yields an infinite steady uniaxial and

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

24

planar extensional viscosities at 1

2 =

, as can be deduced from Eqs. 28–29 resulting from this

model for steady uniaxial, E,U, and planar, E,P, extensional viscosities:

( )( )

0E,U

3

1 2 1

=

− + (28)

( )( )

0E,P

4

1 2 1 2

=

− + (29)

The model also unrealistically predicts a constant steady shear viscosity (equal to 0) and the

second normal stress difference, N2, equal to zero.

The use of the UCM model made it possible to reveal the qualitative role of melt

elasticity in the film casting process. It was found that the edge bead defect was more

pronounced in the Newtonian case than in the viscoelastic case (see Figure 14) and that

increasing the melt elasticity (i.e. the Deborah number) reduces NI [28], which is in good

agreement with the experimental results. This constitutive equation was utilized in the

following studies: 1.D (2 works) [85], [83]; 1.5D (8 works) [27], [89], [53], [94], [99], [29],

[98], [44]; 2D (4 works) [113], [28], [8], [106]; 3D (0).

3.2.4 Generalized Upper-Convected Maxwell model

The generalized Upper-Convected Maxwell model is simply the UCM model, in which the

relaxation time and the shear viscosity are allowed to vary with the second invariant of the

deformation rate tensor:

( ) ( )D DII 2 II D

+ = (30)

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

25

For the film casting modeling, the shear viscosity, ( )DII , was chosen as the Carreau function

(Eq. 24 with = 0) and the relaxation time, ( )DII , in the form of the Eq. 31.

( )

( )0

D n2

t D

II

1 II

=

+

(31)

In this equation, 0, t and n are adjustable parameters. The model can represent steady shear

viscosity, N1 and extensional rheology more realistically than the UCM model but it still shares

the key disadvantages of the original model, i.e., N2 is predicted to be zero and steady uniaxial

and planar extensional viscosities becomes infinite, if the extensional strain rate becomes equal

to ( )1

2 DII. This model was used in the film casting modeling (1.D (1 work in total) [83]) to

understand the role of the power-law exponent n and the characteristic relaxation time on the

velocity profile and the relationship between DR and tensile force.

3.2.5 Giesekus model

The Giesekus model was proposed in 1966 from the simple dumbbell theory for dilute solutions

considering anisotropic drag [119, 123–125]. The model is given as follows:

p s

= + (32)

s

2 D = (33)

2

1 1 pp p pp

2 D

+ + =

(34)

0 p = + (35)

0 1G = (36)

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

26

2

0

1

=

(37)

where is the extra-stress tensor, p

and s

is the polymer and the solvent contribution to the

stress tensor, is the solvent viscosity, p is the polymer viscosity, D is the deformation rate

tensor, 1 is the relaxation time, the symbol represents the upper-convected time derivative,

0 is the zero-shear-rate viscosity, G is the modulus and is the parameter characterizing

anisotropic hydrodynamic drag. The minimum and maximum anisotropies correspond to = 0

and = 1, respectively [125], but as shown by Bird [119], the model behaves realistically only

if 0.5. This model can represent a steady shear viscosity of polymer melts in very wide

shear rate range and it correctly predicts non-zero values of N1 as well as negative value of N2.

On the other hand, its behavior in an extensional flow is not realistic. The key disadvantage of

this model is the inability to predict the decrease in extensional viscosity, if the extensional

strain rate increases. The model has been found to provide reasonable predictions for the film

neck-in, the centerline velocity profile and the temperature drop in the air-gap and it also

predicts an increase in film neck-in and the temperature drop in the air-gap as the air-gap length

is increased [107].

This constitutive equation has been used in the following studies: 1.D (1 work) [79]; 1.5D (0) ;

2D (3 works) [113], [107], [111]; 3D (0).

3.2.6 Modified Giesekus model

The original Giesekus model is not able to realistically represent extensional flows because the

polymeric chains are assumed to be infinitely extensible. In order to overcome this model

drawback, Wiest [126] modified the original Giesekus model [124] by incorporating the finite

extensibility of polymer molecules by using the Peterlin approximation. The set of equations

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

27

remains the same as in the case of the original Giesekus model except for Eq. 34, which must

be replaced by the following expression

( ) ( )21

1 1 pp p pp

Dln Z Dln ZZ 1Z 2 D

Dt 2 Dt

− + + = +

(38)

where is the Kronecker delta and Z is the function defined as

( )p

c

c

tr1Z b 3

b G

= + +

(39)

The term D()/Dt represents substantial time derivative, which is defined as

( ) ( )

( )D

vDt t

= +

(40)

In this model bc represents the chain extensibility parameter. Note that if → cb , →Z 1 , the

modified model is reduced to the original Giesekus model whereas for = 0, the “FENE-P”

dumbbell constitutive equation is recovered [126]. This model shares the same advantages with

the original Giesekus model with the additional ability to describe extensional strain hardening

as well as extensional strain thinning for steady uniaxial as well we as planar extensional

viscosities. In addition, the model is derived from kinetic theory, which allows to relate model

parameters with molecular characteristics. On the other hand, the model predicts that both

uniaxial and planar extensional viscosities are very similar at high deformation rates [84], which

might not be realistic, as indicated in [127].

The model was used to investigate the role of the uniaxial extensional strain hardening

on velocity and stress profiles and film tension in single layer as well as the multilayer film

casting process by using 1.D flow kinematics [84], [75], [76].

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

28

3.2.7 eXtended Pom-Pom model

The eXtended Pom-Pom (XPP) model was proposed by Verbeeten et al. [128]. This model

represents an approximation of the original Pom–Pom model proposed by McLeish and Larson

[129], which is based on the Doi-Edwards reptation tube theory and uses the Giesekus

anisotropy parameter α. The model considers a simplified H topology of branched molecules

and the relaxation time is expressed as a tensor to separate the stretch and orientation. The

model is given by the following set of equations:

( )1

2GD

− + = (41)

where the relaxation time tensor is defined as

( ) ( ) ( )1 1 1 1

0b

1f G f 1

G

− − − − = + + − (42)

Extra function:

( )( )2

1 0b

2 2

s

tr2 1 1f 1 1

3G

− = − + −

(43)

Backbone stretch and stretch relaxation time:

( )tr

13G

= + ,

( )1

s 0se− −

= , 2

q = (44)

where q (number of arms), λ0S (stretch relaxation time) are adjustable parameters, which are

allowed to vary with the orientation relaxation time, λ0b. Note that the Maxwell parameters are

G and λ0b = λ. The model has an excellent capability to describe the shear and extensional

rheology for long-chain branched polymers such as LDPE, which is widely used in the film

casting technology. The model also predicts non-zero values of N1 and N2 as it should be.

Additionally, the model parameters are directly related to the molecular characteristics because

the model is derived from molecular arguments. On the other hand, the model is not suitable

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

29

for linear polymers due to the assumed structural topology and uses a very high number of

adjustable parameters, which makes it difficult to identify them from the measured

experimental data. The model (similarly to the original Pom-Pom model) also predicts that

steady uniaxial and planar extensional viscosities becomes practically identical at high

extensional strain rates, which might not be realistic for some LDPEs, as indicated in [127] (see

Figure 15). The XPP model correctly predicts that increasing DR and the air-gap reduce NI.

According to the authors, the agreement between the experimental data (LDPE [41], [42] and

long chain branched PP [43]) and the simulation results was qualitative rather than quantitative

in terms of necking behavior (see Figure 16 as the example for LDPE; here the 8 mode XPP

model, in which all model parameters were allowed to vary with the relaxation mode, was used;

step shear and step uniaxial extensional experiments were used to validate the XPP model).

This constitutive equation was utilized in the following studies: 1.D (0); 1.5D (3 works) [41],

[42], [43]; 2D (0); 3D (0).

3.2.8 Rolie–Poly Stretch (RP-S) model

Rolie–Poly (ROuse LInear Entangled POLYmer) stretch model is a tube-based model proposed

by Likhtman and Graham [130], which takes into account the convective constraint release,

reptation and chain retraction. The model is given as

( )( )

( )0

c

d r

2 1 3 tr tr1

3

= − − − + −

(45)

where d and

r are the reptation and the Rouse relaxation times, respectively; c is the

convective constraint release coefficient, 0 is a fitting scalar parameter. Being c equal to

zero, 0 parameter does not have to be specified. This model has showed a good capability to

describe the transient shear and extensional rheology of linear film casting resins (namely

LLDPE and HDPE), it is mathematically simple and gives a non-zero value of N1. On the other

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

30

hand, the model is not able to describe the rheological behavior of branched polymers and

unrealistically predicts N2 = 0 [131]. This constitutive equation was utilized in the following

theoretical studies summarized in this review: 1.D (0); 1.5D (3 works) [41], [42], [43]; 2D (0)

; 3D (0). The model predicted an increase in NI with increased DR and the air-gap. According

to the authors, the agreement between the experimental data (HDPE [41]; LLDPE [41], [42];

and linear PP [43]) and the simulation results was qualitative rather than quantitative with

respect to NI (see Figure 17 as an example for LLDPE; here the 8 mode RP-S model, in which

d and r were allowed to vary with the relaxation mode, was used; = 0 and 0 0.5 = − ; step

shear and step uniaxial extensional experiments were used to validate the RP-S model).

3.2.9 Modified Leonov model

The modified Leonov model is based on heuristic thermodynamic arguments resulting from the

theory of rubber elasticity [132–137]. In this constitutive equation, a fading memory of the melt

is determined by an irreversible dissipation process driven by the dissipation term, b . This

model relates the stress and elastic strain stored in the polymer melt as:

1

c c

W W2 c c

I II

= −

(46)

where W is the elastic potential, which depends on the invariants cI and cII of the recoverable

Finger tensor c ,

( )

( )L Ln 1 n 1

c c

L

I II3GW 1 1 1

2 n 1 3 3

+ + = − − + −

+

(47)

where G denotes the linear Hookean elastic modulus, and Ln are numerical parameters.

Leonov assumed that the dissipative process acts to produce an irreversible rate of strain p

e

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

31

c c1

p

I IIe b c b c

3 3

= − − −

(48)

which spontaneously reduces the rate of elastic strain accumulation. Here, δ is the unit tensor

and b stands for the dissipation function defined by Eq. 50. This elastic strain c is related to

the deformation rate tensor D as follows

p

c c D D c 2c e 0− − + =

(49)

where c is the Jaumann (corotational) time derivative of the recoverable Finger strain

tensor. The dissipation function b proposed in [64] is given as

( )( )

( )c

c c

c

sinh I 31b I exp I 3

4 I 3 1

− = − − + − +

(50)

where and are adjustable parameters of the model.

The model has a very good capability to describe shear viscosity, N1, N2, uniaxial and planar

extensional viscosities for linear as well as branched polymers [57, 58, 62–64, 138]. The model

also offers an independent control of uniaxial and planar extensional viscosities, which was

used for systematic investigation of the role of planar to uniaxial extensional viscosity ratio on

the film casting process for different LDPEs. On the other hand, the interpretation of molecular

meaning of the used model parameters is limited because the model is derived from

thermodynamics rather than molecular arguments. Note that the original Leonov model is

recovered if nL = = = = 0.

The model demonstrated the ability to describe film casting experimental data for linear (PP)

as well as branched (LDPE) polymers even by using a single mode (i.e. utilizing a single pair

of and G) only [35, 37, 38] (see Figure 18; In the case of PP, the lowest relaxation time typical

for polyolefins was chosen and G was calculated from the known 0. The nonlinear model

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

32

parameters were adjusted as the typical values for linear polymers, i.e. = 0, nL = 0, = 0.5,

= 0.5. In the case of LDPE, the model parameters, namely , G, , , were identified using

deformation rate dependent ‘steady state’ uniaxial extensional viscosity experimental data and

known 0. The parameter nL was kept equal to zero). The modified Leonov model was used in

the following studies: 1.D (0); 1.5D (3 works) [37], [38], [35]; 2D (0); 3D (0). The original

Leonov model was used only in: 1.5D (1 work) [25].

3.2.10 Phan-Thien-Tanner (PPT) model

The model was derived by Phan-Thien and Tanner [139] and Phan-Thien [140] from the Lodge-

Yamamoto network theory, in which junctions are allowed to form and decay due to the flow.

The model is given by the following equation

( ) 02 D

+ = (51)

This model uses the Gordon–Schowalter convected time derivative of the stress tensor,

,

defined as

( )T

vt

= + − −

,

pvL D= − (52)

and the relaxation time, ( ) , which is allowed to vary with the trace of the stress linearly

(Eq. 53) or exponentially (Eq. 54).

Linear PTT model:

( ) ( )p1 tr

G

= + (53)

Exponential PTT model:

( ) ( )pexp tr

G

=

(54)

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

33

The model utilizes two parameters, p and p (together with and G, where 0 = G). The linear

PTT predicts an unrealistically monotonically increasing extensional viscosity, while the

exponential PTT has the ability to give a maximum in the extensional viscosity, when plotted

as a function of the extensional strain rate. Thus, the exponential PTT model is used more in

modeling of the film casting process than its linear version; it has a good ability to describe

shear as well as extensional rheology of linear as well as branched polymers and predicts non-

zero values for N1 and N2. On the other hand, steady uniaxial and planar extensional viscosities

are predicted to be practically identical at high extensional strain rates, which might not be

realistic for some polymers, as indicated in [127] (see Figure 19). This model has been shown

to provide good agreement with experimental data for LDPE over a wide range of the take-up

velocity and the air-gap length [109], [54] (see Figure 20). In this case, the exponential (6-7)

mode PTT model using p and p model parameters varying with each relaxation mode was

necessary to adequately represent the measured shear and extensional data LDPEs reasonably.

The discrete relaxation spectrum was determined from frequency-dependent loss and storage

moduli, while p and p parameters were identified from steady state shear viscosity vs. shear

rate and steady uniaxial extensional viscosity vs. extensional strain rate.

This constitutive equation was utilized in the following studies within this review:

1.D (1 work) [79]; 1.5D (2 works) [47], [44]; 2D (7 works) [46], [105], [36], [108], [109], [54],

[40]; 3D (1 work) [116]. The linear variant of the PTT model was used only in: 2D (1 work)

[26].

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

34

3.2.11 Larson model

This model represents a differential approximation of the Doi-Edwards (DE) integral tube

model proposed by Larson [141, 142], which includes non-affine motion and is given by the

following expression:

( ) 0

2D : G 2 D

3 G

+ + + = (55)

where is the only non-linear model parameter (note that the best approximation of the DE

model is achieved, if 3 5 = [125]). The behavior of the model is comparable to the

exponential PTT model meaning that it can capture shear and extensional viscosities for both

linear and branched polymers, and N1 is predicted to be a non-zero value. However, the model

unrealistically predicts N2 = 0, steady uniaxial and planar extensional viscosities becomes

comparable at high extensional strain rates, and must be varied with the relaxation time to

capture uniaxial extensional strain hardening for commercial types of highly branched

polyethylenes over a wide range of extensional strain rates. This model was successfully used

in the film casting modeling for 4 different types of LDPEs, but 13 pairs of , G, were needed

to describe the experimental reality [39]. The model provides reasonably good predictions for

the neck-in and the film thickness (see Figure 21). This model also predicts that the increase in

the uniaxial extensional strain hardening firstly, decreases NI in agreement with experimental

observations (see Figure 22), and secondly, the edge beading region becomes narrower.

This constitutive equation was utilized in the following studies: 1.D (0); 1.5D (0) ; 2D (2

works) [39], [105]; 3D (0).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

35

3.2.12 Integral constitutive equation of the K-BKZ type with PSM strain-memory

function

The multi-mode Kaye-Bernstein-Kearsley-Zapas (K-BKZ) type of the integral model

proposed by Papanastasiou, Scriven, Macosko (PSM) [143] and then modified by Luo and

Tanner [144] to account for N2 through the parameter has the following expression:

( )( ) ( )

1 1

Nt 1j j

j 1 j j j j jC C

G1 t texp C t C t dt

1 3 I 1 II− −

−=

− = − + − − + + −

(56)

where t and t are the times present and past, respectively, 1−C

I and 1−C

II are the first and

second invariant of the Finger strain tensor 1−

C ; j , j and are the parameters of the model.

The K-BKZ model reduces to the upper-convected Maxwell model when = 0 and → + .

The model has a very good ability to describe the shear viscosity, N1, N2 and uniaxial

extensional viscosities for linear and branched polymers. The main disadvantage of the model

is the inability to predict extensional strain hardening in planar extensional viscosity for

branched polymers as LDPE [145] (see Figure 23).

The model showed a good agreement between experimental data and previous film casting

simulations based on the Newtonian and UCM results in terms of film thickness, width and

temperature [89]. Note that in order to capture the shear and extensional rheology of LDPE

melts, the parameter in Eq. 56 needs to be varied with the relaxation time, otherwise the

extensional strain hardening is overpredicted at high extensional strain rates (see Figure 23).

This constitutive equation was utilized in the following studies: 1.D (1 work) [45];

1.5D (1 work) [89]; 2D (0); 3D (0).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

36

3.2.13 Hooke’s Law with creep

Smith and Stolle considered the viscoelastic polymer melt to be an elastic material that is

creeping. For such a description, they have suggested the incremental form of Hooke’s law of

the following form [146]:

( )cE = − (57)

where is the stress increment, E is the elasticity matrix, is the total strain increment

and c

is the modified Perzyna creep strain increment that depends on the time step, the

effective creep strain rate and the second invariant of the deviatoric stress tensor. This model

was used to theoretically investigate the role of elasticity (by changing the relaxation time) and

extensional strain hardening. It was found that the predictions were consistent with predictions

of the conventional viscoelastic model i.e. that NI decreases with increased extensional strain

hardening and the relaxation time. Note that the model uses only one constant relaxation time

and modulus, and its ability to describe rheological experimental data has not been provided by

the authors.

This constitutive equation was utilized in the following studies: 1.D (2 works) [77], [78];

1.5D (0) ; 2D (1 work) [103]; 3D (0).

3.3 Boundary conditions for viscoelastic constitutive equations

If viscoelastic constitutive equations are used, the additional boundary stress condition at

the die exit must be specified. This boundary condition is given by both, the flow in the die

(upstream) and the extensional flow in the drawing length (downstream). Accurate

determination of this additional boundary stress value therefore requires an intensive numerical

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

37

computation [46]. The following paragraphs summarize the approaches used to determine these

types of boundary conditions (see also Tables 2–5).

Anturkar and Co [83] in their study, using the generalized upper-convected Maxwell model,

estimated axial component of the stress tensor, xx , as the mean stress value for a

fully-developed slit flow in an infinitely wide die. Silagy et al. [27] and [28] based on Denn et

al. [147] and Demay et al. [88, 148] assumed two different stress states at the end of the die. In

the first case, the extra stress in the machine direction, xx , equals zero, so the extra stresses are

completely released due to the die swell, or the second, which assumed the mean value of the

extra stress after flow in an infinite die with a rectangular cross section while the transversal

extra stress, yy , is set to the value obtained from the Newtonian solution. They found that the

initial stress conditions at the die exit had little effect on the final shape of the film, but

calculations were performed only for low Deborah numbers. Iyengar and Co [84] took a

different approach, and instead of specifying the axial stress component, set the ratio zz xx to

the value between two extreme cases for planar extensional flow and a fully-developed slit flow

in the die noting that the true stress ratio should have lied in their range. Iyengar [149] then

reported that both extreme cases with corresponding stress ratios provide very similar velocity

and stress profiles. Debbaut at al. [113] in their viscoelastic study assumed an initial stresses of

zero. Same as in Smith’s work [8].

For a multilayer film casting analysis (based on a single-mode modified Giesekus model)

Pis-Lopez and Co [76, 150] showed that if the aspect ratio (defined here as the ratio of the total

film thickness at the die exit to the drawing distance 2e0/X) is smaller than 0.05, the velocity

and stress profiles converge to the same values, regardless of whether the initial stress condition

is based on the assumption of a fully-developed slit flow or a fully-developed planar extensional

flow. In another study using a multi-mode model approach, Denn [151] left the longest

relaxation mode unspecified at the die exit, and rest of the modes were set with respect to this

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

38

mode. In contrast, Christodoulou et al. [46] concluded that the shortest mode should be left

unspecified with the reasoning that the longest mode p

xx( N ) is mainly determined by the flow

inside the die, while the shorter modes p

xx( j ) are determined by the external flow in the air gap.

Beris and Liu [152, 153] in their fiber spinning study for the single mode UCM viscoelastic

liquid specified the die exit stress state through the ratio of normal to axial stress, yy xx , and

not each component separately. This value was approximated as the value bellow the

homogeneous steady extensional flow at the effective extensional strain rate. For the

viscoelastic multimode model, Denn [151] specified also p p

xx( j ) xx( N ) for j < N as an extra

condition to p p

yy( j ) xx( j ) for all relaxation modes.

Devereux and Denn [154] proposed the same distribution between partial stresses as in the

case of a fully-developed capillary flow with neglected radial partial stresses. The remaining

initial stresses were adjusted to satisfy the downstream boundary condition (see Eq. 58).

p

xx( j) j j

Np

xx(N)j j

j 1

G

G=

=

(58)

Note that Gagon and Denn [155] simplified the above-mentioned relationship for the wedge

spectrum into

xx( j) j

xx(N) N

=

(59)

Barborik and Zatloukal [37, 38] recently defined the state of the stress at the die exit by the

ratio of the second to the first normal stress difference, −N2/N1, calculated from the upstream

side by using the viscoelastic modified Leonov constitutive equation.

( ) ( )

( ) ( )

zz yy2

1 xx zz

0 0N

N 0 0

− − = −

− (60)

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

39

It has been found that if the Deborah number is less than 0.1, the choice of initial stress

conditions at the die exit has little effect on the steady-state calculations. However, at higher

Deborah numbers, the die exit stress state, which may be influenced, for example, by extrusion

die design, starts to significantly affect the neck-in phenomenon.

3.4 Overview on steady-state film casting modeling

Initial efforts have been made for a fiber spinning process in which the flow kinematics are

mathematically similar when considered as a one-dimensional flow, for the Newtonian and

Maxwell fluids by Gelder [156] and Fisher [157, 158], respectively. These studies focused on

the draw resonance phenomenon, that Christensen [159] and Miller [160] first encountered, and

who postulated that the nature of this phenomenon is not of viscoelastic, as it can also be

observed in Newtonian fluids. Extending the kinematics of the process to two or three

dimensions, both processes become different and one can observe phenomena in the film

casting, which have no counterpart in the fiber spinning, i.e. neck-in and edge-beading. The

above preliminary studies provided the background for extended studies on the extrusion film

casting (EFC). Initial attempts to simulate EFC operations were devoted to investigating the

stability of the process and determining the onset of draw resonance, rather than quantifying

the extent of the neck-in phenomenon. The very first study of EFC process modeling in this

way was carried out by Yeow [80] using numerical modeling. For the steady state solution,

he used a one-dimensional isothermal model for Newtonian fluid (planar extensional free

surface flow) and investigated the effect of introduced small two-dimensional perturbances on

flow stability (namely transverse perturbations). The edge-effects, surface tension,

aerodynamic drag and fluid inertia and gravity were neglected. A small curvature of the film

was assumed along with a uniform axial stress and axial velocity over the film thickness. Due

to the assumed free surface flow kinematics in the drawing section, the model could not capture

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

40

the edge-bead defect and the shrinkage in the width of the film, which was considered to be

infinitely wide. The film thickness could only change in the machine direction.

Aird and Yeow [81] continued this mathematical background for the 1-dimensional

(1D) model and extended the analysis for power-law fluids. Consequently, Anturkar and Co

[83] and Iyengar and Co [75, 84] used isothermal generalized Upper-Convected Maxwell and

Giesekus constitutive equations for linear and non-linear analysis in viscoelastic fluid

simulations. The first isothermal attempts to model the necking phenomena were performed by

Sergent [161] in 1977 and then by Cotto, Duffo and Barq [18, 87, 88] for non-isothermal

conditions.

Another milestone work was set by Dobroth and Erwin [55] in 1986, who pointed out that

there is a planar and uniaxial extensional deformation at the center of the film and at the edge,

respectively, and that the extent of the edge-beads and the interrelated neck-in phenomenon is

determined by the interplay between these two regions through an edge stress effect. This

idea was consequently confirmed experimentally by Ito et al. [52]. The assumption of a planar-

uniaxial distribution in the Dobroth and Erwin model can therefore be considered reasonable.

Just note that, in the case of fiber spinning, only a uniaxial extensional flow can be observed.

Some authors have attempted to relate and quantify the gauge of the observed necking in

terms of rheological parameters, such as shear, uniaxial and planar extensional viscosity. Many

authors have reported that the strain hardening in uniaxial extension may reduce the extent of

the necking phenomena [39, 72, 74, 113]. This idea was continued by Ito [53], who related the

neck-in intensity to rheological parameters, such as the ratio of planar viscosities in axial and

transverse directions, and derived an analytical equation for the edge line of Newtonian and

Maxwell fluids. In accordance with the article by Dobroth and Erwin [55], who, as the first

recognized deformation type in the drawing area, Shiromoto [7, 8, 33] recently presented the

idea that the extent of the necking should not have been described by uniaxial extensional

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

41

viscosity only in addition to take-up length but as the ratio of planar and uniaxial extensional

viscosities reflecting the deformation type in the central and edge portion of the film in the

drawing section. In addition to performing non-isothermal viscoelastic simulations, they also

proposed a theoretical model based on the force balance and the deformation type of the film

in order to predict necking behavior.

More recently, the 2-dimensional (2D) membrane model has been introduced by d’Halewyu

[112] and Debbaut [113] for Newtonian and viscoelastic fluids, respectively. This frequently

used model is able to predict the dog-bone defect, i.e. development of edge-beads under

stationary conditions.

Silagy et al. [27] improved and enriched the membrane model with a supplementary

kinematic hypothesis originally provided by Narayanaswamy [162] in his paper on float glass

stretching, and performed an extended isothermal study on the effect of processing conditions

on film geometry and EFC stability analysis for Newtonian and Upper convected Maxwell

(UCM) fluids. Due to the assumptions used in the flow kinematics, this model was able to cover

the reduction of the film width, and thus predict the neck-in phenomenon, but still could not

predict the edge-beading. This limitation was removed in their subsequent work [28], where a

2D isothermal membrane model combined with the Phan–Thien and Tanner (PTT) constitutive

equation was developed and the obtained steady and transient stability results were compared

with its 1.5D predecessor. In the following years, the 1.5D version of Sylagy’s membrane model

was used in many studies, and much work was done on EFC under non-isothermal conditions

including crystallization effects (Lamberti et al. [91–93, 163], Lamberti and Titomanlio [92,

94, 95, 164], and Lamberti [67]). A three dimensional model for EFC simulation was further

developed by Sakaki et al. [114]. To solve the model equations, it was necessary to use the

finite element method. The problem was considered isothermal, steady state and Newtonian. A

process parameter space was chosen to reflect the industrial processing conditions. The model

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

42

captured the development of both the neck-in and the edge beading and the effect of DR, X and

die width were investigated. They found that the intensity of the neck-in and the edge beading

was affected by DR and X but not by the width of the die. The extent of the neck-in increased

with increasing DR and X. Lately, this approach has been extended by Zheng et al. [115] for a

non-isothermal steady Newtonian fluid. Kometani et al. [111] performed experimental and

theoretical investigations of the effects of rheological properties on the neck-in in the film

casting. For the two tested PP and LDPE materials, without significant differences in

viscoelastic properties, except for extensional ones (LDPE showed a remarkable increase in

extensional viscosity at high strain rates), the neck-in extent for PP under higher draw ratios

increased over LDPE where the neck-in was constant and independent of the draw ratio. Based

on these experiments, the authors concluded that neck-in phenomenon in film casting depends

on the extensional rheological properties. Furthermore, they utilized a simulation based on three

deferent rheological models (the Newtonian, Bird-Carreau and Giesekus model) to assess its

applicability for film casting modeling. The results obtained by simulation based on the

Giesekus model were in quantitative agreement with the experimental observations for both

polymers, but the other two models used did not provide predictions well describing the

measured data due to their inaccurate representation of extensional viscosities.

The influence of macromolecular architecture on the extent of the necking phenomenon has

been investigated by Ito et al. [52, 53] (effects of the draw ratio and the take-up length on the

necking for LDPE, HDPE and mLLDPE) and Baird et al. [68, 69] (effects of long chain

branching and molecular weight distribution on the necking for LDPE, mLLDPE and Ziegler-

Natta catalyzed LLDPE). Research on a multi-layer film casting considering Giesekus fluid has

been performed in studies of Pis-Lopez and Co for steady state [76] and stability analysis [150].

Recently, Pol et al. [41, 42] and Chikhalikar et al. [43] published a number of articles in

which they performed experimental and theoretical investigations of the effects of long chain

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

43

branching and molecular weight distribution on the extent of the necking phenomenon. For this

purpose, they have used the 1.5D membrane model, originally proposed by Silagy [27], the

multi-mode eXtended Pom-Pom constitutive equation and the multi-mode Rolie-Poly stretch

constitutive equation, respectively, for the long chain branched (LDPE, PP) and the linear

(HDPE, PP) polymers. Fixing the DR and X, they found that the extent of the necking was

smaller for HDPE with a broader molecular weight distribution than for LLDPE with a narrower

molecular weight distribution, and further that long chain branched LDPE necks-in to lower

extent than linear HDPE or LLDPE (i.e. that increase in long chain branching is more effective

in neck-in suppression than just broadening MWD). In a subsequent study, Pol and Thete [98]

switched from the 1.5D model that was used in their predecessor works on this topic to the

two-dimensional model originally proposed by Ito et al. [53] incorporating UCM constitutive

equation. In addition, they derived an analytical solution for low and high Deborah numbers.

They found that while the film width of the modelled LLDPE continuously decreased with

increased draw ratio, the film width for LDPE decreased with increased draw ratio for long

take-up lengths and remained constant for shorter ones. This means that there is a locus of

points in the attainable region that divides DR–De plane into sections where the dependence of

the neck-in on the draw ratio has opposite trends.

In their latest work [44], they focused on the effects of the individual viscoelastic relaxation

modes of a polymer melt on its behavior in polymer melt extrusion film casting process using

UCM and PTT constitutive equations and the 1.5D isothermal membrane model. They found

that experimental data for long-chain branched LDPE was better described using the UCM

model, while the PTT model provided better simulation results for the experimental data with

linear LLDPE.

Although the actual EFC manufacturing process involves complex 3-dimensional (3D)

kinematics whose numerical simulation can be very challenging, many authors have shown that

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

44

the EFC 1.5D membrane model (originally proposed by Silagy) is able to provide results that

are in good agreement with experiment data (if appropriate constitutive equations are used).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

45

4 CONCLUSIONS

Modeling of the steady-state film casting process began in 1974 using a 1D kinematic

approach. 1.5D and 2D kinematic models were developed at 1990 and are currently the most

commonly used (in more than 73%). A full 3D kinematic approach was introduced in 1996, but

due to its complexity, its use is very rare (less than 5%) and usually only applies to viscous

fluids. The following most commonly used simplifying assumptions are used to overcome

numerical difficulties in steady-state cast film process modeling: reduction in dimensionality

(1D, 1.5D and 2D), isothermal conditions – constant temperature field, non-transient

description, incompressible fluid, constant density, excluded inertia effects, excluded effects of

gravitational forces, constant boundary conditions, unrealistic or simplified constitutive

equations, neglected aerodynamic drag, neglected surface tension, neglected die swell,

neglected edge-effects, excluded crystallization (temperature, flow-induced), neglected the sag

of the film in non-vertical installations (film curvature), effects of other devices (air knife,

vacuum box, electrostatic pinning).

The following constitutive equations have already been used in the steady-state modeling

of the extrusion cast film process: Viscous – Newtonian model, Generalized Newtonian model

(utilizing Cross, Carreau and Carreau-Yasuda viscosity functions); Elastic – incremental Hooks

law with creep. Viscoelastic – upper convected Maxwell model, generalized upper convected

Maxwell model with deformation rate dependent relaxation time and viscosity, linear and

exponential PTT model, Larson model, K-BKZ model with Papanastasiou-Scriven-Macosko

(PSM) damping function, Leonov model modified by specific type of recoverable strain

dependent dissipation function, Giesekus model, modified Giesekus with finite chain stretch,

eXtended Pom-Pom model (XPP) and RoliePoly-Stretch model (RP-S).

Intensive experimental research on extrusion film casting in relation to the neck-in began in

1986 and continues to this day. Most of the research reported in the open literature focuses on

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

46

polyolefins such as LDPE, LLDPE and PP, while experimental studies for other polymers

(PET, HDPE or polyethersulfone - PES) are very rare.

Based on experimental and theoretical studies presented in the open literature, the following

variables were found to be the key with respect to the neck-in phenomenon: draw ratio, Deborah

number (i.e. melt relaxation time, melt velocity at the die exit, air-gap), film cooling, ratio of

the second and the first normal stress difference at the die exit, uniaxial extensional strain

hardening and planar-to-uniaxial extensional viscosity ratio. Uniaxial and planar extensional

viscosities can be considered as key material parameters, because during the polymer melt

stretching in the post die area, the middle of the film undergoes planar elongation, while the

material at the edge undergoes uniaxial elongation. However, measuring uniaxial and planar

extensional viscosities at very wide deformation rates is still a very difficult task because the

generation and control of extensional flows is difficult. Likewise, the role of draw ratio, heat

transfer coefficients, flow induced crystallization, flow history (generated inside the extrusion

die as –N2/N1 ratio), molecular architecture of polymer melts on the film formation and its

stability is still not fully understood yet due to the absence of relevant experimental and

theoretical studies.

There is a currently attempt to relate flow stability with the molecular characteristics of the

polymers used through advanced molecular constitutive equations such as XPP and RP-S.

However, their use is still limited because these models do not allow realistic rheological

description of polymers or polymer mixtures with an internal structure, which does not

correspond to the molecular assumptions under which these constitutional equations were

derived. Another major problem in experimental and theoretical flow stability research is the

neglect of the influence of memory and flow history of polymer melts in extrusion heads on the

stability of film formation in the post die area, as evidenced by recently published work.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

47

Further research in the following areas can be expected to help significantly understand the

formation stability of polymeric flat films:

• development and use of suitable rheological techniques for the determination of

uniaxial and especially planar extensional viscosities at high deformation rates,

• development/modification and use of advanced constitutive equations with ability to

handle measured extensional rheology data for polymers, polymer blends and filled

polymers,

• elucidation of the role of polymer molecular architecture, flow history, heat transfer

coefficients and flow induced crystallization,

• understanding the role of the draw ratio in film formation for branched polymers.

Data availability statement

Data sharing is not applicable to this review article as no new experimental data were generated.

Acknowledgments

The authors would like to acknowledge the Institutional Support Project 2020 (Polymer Centre

at Faculty of Technology, Tomas Bata University in Zlin).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

48

5 REFERENCES

1. R. W. Baker, “Membrane Technology and Applications,” (John Wiley & Sons, Ltd.,

2012).

2. P. Arora, and Z. (John) Zhang, “Battery Separators,” Chem. Rev. 104 (10), 4419 (2004).

3. S. S. Zhang, “A review on the separators of liquid electrolyte Li-ion batteries,” J. Power

Sources 164 (1), 351 (2007).

4. M. Ulbricht, “Advanced functional polymer membranes,” Polymer 47 (7), 2217 (2006).

5. T. Kanai, and G. A. Campbell, “Film Processing,” (Hanser Publishers, Munich, 1999).

6. T. Kanai, and G. A. Campbell, “Film Processing Advances,” (Hanser Publishers,

Munich, 2014).

7. K. Christodoulou, S. G. Hatzikiriakos, and E. Mitsoulis, “Numerical simulation of the

wire-pinning process in PET film casting: Steady-state results,” AIChE J. 58 (7), 1979

(2012).

8. W. S. Smith, “Simulating the cast film process using an updated Lagrangian finite

element algorithm,” Ph.D. thesis, McMaster University, 2001.

9. K. Resch, G. M. Wallner, C. Teichert, and M. Gahleitner, “Highly transparent

polypropylene cast films: Relationships between optical properties, additives, and

surface structure,” Polym. Eng. Sci. 47 (7), 1021 (2007).

10. Z. Tadmor, and C. G. Gogos, “Principles of Polymer Processing, 2nd Edition,” (John

Wiley & Sons, Hoboken, New Jersey, 2006).

11. J. R. A. Pearson, “Mechanics of Polymer Processing,” (Elsevier Applied Science

Publishers, London, 1985).

12. W. S. Smith, “Nonisothermal film casting of a viscous fluid,” Ph.D. thesis, McMaster

University, 1997.

13. M. Zatloukal, and R. Kolarik, “Investigation of convective heat transfer in 9-layer film

blowing process by using variational principles,” Int. J. Heat Mass Transf. 86, 258

(2015).

14. R. Kolarik, M. Zatloukal, and M. Martyn, “The effect of polyolefin extensional rheology

on non-isothermal film blowing process stability,” Int. J. Heat Mass Transf. 56 (1–2),

694 (2013).

15. R. Kolarik, and M. Zatloukal, “Modeling of nonisothermal film blowing process for non-

Newtonian fluids by using variational principles,” J. Appl. Polym. Sci. 122 (4), 2807

(2011).

16. M. Zatloukal, and J. Vlcek, “Application of variational principles in modeling of the film

blowing process for high stalk bubbles,” J. Nonnewton. Fluid Mech. 133 (1), 63 (2006).

17. M. Zatloukal, and J. Vlcek, “Modeling of the film blowing process by using variational

principles,” J. Nonnewton. Fluid Mech. 123 (2–3), 201 (2004).

18. D. Cotto, P. Duffo, and J. M. Haudin, “Cast film extrusion of polypropylene films,” Int.

Polym. Process. 4 (2), 103 (1989).

19. S. A. Ashter, “Introduction to Bioplastics Engineering,” Plastics Design Library

(William Andrew Publishing, Oxford, 2016).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

49

20. F. Wu, M. Misra, and A. Mohanty, in Annual Technical Conference - ANTEC,

Conference Proceedings 2019-March, (Society of Plastics Engineers, Detroit; United

States, 2019).

21. T. I. Burghelea, H. J. Grieß, and H. Münstedt, “An in situ investigation of the draw

resonance phenomenon in film casting of a polypropylene melt,” J. Nonnewton. Fluid

Mech. 173–174, 87 (2012).

22. P. Barq, J. M. Haudin, J. F. Agassant, H. Roth, and P. Bourgin, “Instability phenomena

in film casting process,” Int. Polym. Proc. 5 (4), 264 (1990).

23. Y. Demay, and J. F. Agassant, “An Overview of Molten Polymer Drawing Instabilities,”

Int. Polym. Process. 29 (1), 128 (2014).

24. R. G. Larson, “Instabilities in viscoelastic flows,” Rheol. Acta 31 (3), 213 (1992).

25. Y. Kwon, “One-dimensional modeling of flat sheet casting or rectangular fiber spinning

process and the effect of normal stresses,” Korea-Australia Rheol. J. 11 (3), 225 (1999).

26. N. D. Polychronopoulos, and T. D. Papathanasiou, “A study on the effect of drawing on

extrudate swell in film casting,” Appl. Rheol. 25 (4), 1 (2015).

27. D. Silagy, Y. Demay, and J. F. Agassant, “Study of the stability of the film casting

process,” Polym. Eng. Sci. 36 (21), 2614 (1996).

28. D. Silagy, Y. Demay, and J. F. Agassant, “Stationary and stability analysis of the film

casting process,” J. Nonnewton. Fluid Mech. 79 (2–3), 563 (1998).

29. S. Bourrigaud, G. Marin, V. Dabas, C. Dupuy, and D. Silagy, “The draw ratio–Deborah

number diagram: A useful tool for coating applications,” Polym. Eng. Sci. 46 (3), 372

(2006).

30. C. W. Seay, and D. G. Baird, “Sparse Long-chain Branching’s Effect on the Film-casting

Behavior of PE,” Int. Polym. Process. 24 (1), 41 (2009).

31. H. J. Grieß, “Fließverhalten von Polypropylenschmelzen innerhalb und außerhalb einer

Kleiderbügeldüse (Flow properties of polypropylene melts inside and outside a coat-

hanger die),” Ph.D. thesis, University Erlangen-Nuremberg, 2014.

32. H. Münstedt, “Extensional rheology and processing of polymeric materials,” Int. Polym.

Process. 33 (5), 594 (2018).

33. H. Münstedt, “Elastic Behavior of Polymer Melts, Rheology and Processing,” (Hanser

Publications, Munich, Cincinnati, 2019).

34. H. Münstedt, in AIP Conference Proceedings (eds. Zatloukal, M. & Musil, J.) 2107,

030001 (American Institute of Physics Inc., Zlin, Czech Republic, 2019).

35. T. Barborik, and M. Zatloukal, “Effect of Heat Transfer Coefficient, Draw Ratio and Die

Exit Temperature on the Production of Flat Polypropylene Membranes,” Phys. Fluids 31

(5), 053101 (2019).

36. D. M. Shin, J. S. Lee, J. M. Kim, H. W. Jung, and J. C. Hyun, “Transient and steady-

state solutions of 2D viscoelastic nonisothermal simulation model of film casting process

via finite element method,” J. Rheol. 51 (3), 393 (2007).

37. T. Barborik, M. Zatloukal, and C. Tzoganakis, “On the role of extensional rheology and

Deborah number on the neck-in phenomenon during flat film casting,” Int. J. Heat Mass

Transf. 111, 1296 (2017).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

50

38. T. Barborik, and M. Zatloukal, “Effect of die exit stress state, Deborah number, uniaxial

and planar extensional rheology on the neck-in phenomenon in polymeric flat film

production,” J. Nonnewton. Fluid Mech. 255, 39 (2018).

39. N. Satoh, H. Tomiyama, and T. Kajiwara, “Viscoelastic simulation of film casting

process for a polymer melt,” Polym. Eng. Sci. 41 (9), 1564 (2001).

40. S. Shiromoto, “The Mechanism of Neck-in Phenomenon in Film Casting Process,” Int.

Polym. Process. 29 (2), 197 (2014).

41. H. V. Pol, S. S. Thete, P. Doshi, and A. K. Lele, “Necking in extrusion film casting: The

role of macromolecular architecture,” J. Rheol. 57 (2), 559 (2013).

42. H. V. Pol, S. Banik, L. B. Azad, S. S. Thete, P. Doshi, and A. Lele, “Nonisothermal

analysis of extrusion film casting process using molecular constitutive equations,”

Rheol. Acta 53 (1), 85 (2014).

43. K. Chikhalikar, S. Banik, L. B. Azad, K. Jadhav, S. Mahajan, Z. Ahmad, S. Kulkarni, S.

Gupta, P. Doshi, H. V. Pol, and A. Lele, “Extrusion film casting of long chain branched

polypropylene,” Polym. Eng. Sci. 55 (9), 1977 (2015).

44. S. S. Thete, P. Doshi, and H. V. Pol, “New insights into the use of multi-mode

phenomenological constitutive equations to model extrusion film casting process,” J.

Plast. Film Sheeting 33 (1), 35 (2017).

45. S. M. Alaie, and T. C. Papanastasiou, “Film casting of viscoelastic liquid,” Polym. Eng.

Sci. 31 (2), 67 (1991).

46. K. Christodoulou, S. G. Hatzikiriakos, and D. Vlassopoulos, “Stability analysis of film

casting for PET resins using a multimode Phan-Thien-Tanner constitutive equation,” J.

Plast. Film Sheeting 16 (4), 312 (2000).

47. R. Dhadwal, S. Banik, P. Doshi, and H. V. Pol, “Effect of viscoelastic relaxation modes

on stability of extrusion film casting process modeled using multi-mode Phan-Thien-

Tanner constitutive equation,” Appl. Math. Model. 47, 487 (2017).

48. K. Canning, and A. Co, “Edge effects in film casting of molten polymers,” J. Plast. Film

Sheeting 16 (3), 188 (2000).

49. K. Frey, C. D. Cleaver, and G. Jerdee, in Proceedings of the 2000 TAPPI Polymers,

Laminations, and Coating Conference (2), 709 (Chicago, IL; United States, 2000).

50. J. S. Lee, H. W. Jung, and J. C. Hyun, in Annual Technical Conference - ANTEC,

Conference Proceedings 1, 54 (2003).

51. J. S. Lee, H. W. Jung, and J. C. Hyun, “Stabilization of film casting by an encapsulation

extrusion method,” J. Nonnewton. Fluid Mech. 117 (2–3), 109 (2004).

52. H. Ito, M. Doi, T. Isaki, M. Takeo, and K. Yagi, “2D Flow Analysis of Film Casting

Process,” J. Soc. Rheol. Japan 31 (3), 149 (2003).

53. H. Ito, M. Doi, T. Isaki, and M. Takeo, “A Model of Neck-in Phenomenon in Film

Casting Process,” J. Soc. Rheol. Japan 31 (3), 157 (2003).

54. S. Shiromoto, Y. Masutani, M. Tsutsubuchi, Y. Togawa, and T. Kajiwara, “The effect

of viscoelasticity on the extrusion drawing in film-casting process,” Rheol. Acta 49 (7),

757 (2010).

55. T. Dobroth, and L. Erwin, “Causes of edge beads in cast films,” Polym. Eng. Sci. 26 (7),

462 (1986).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

51

56. F. N. Cogswell, “Measuring the Extensional Rheology of Polymer Melts,” Trans Soc

Rheol 16 (3), 383 (1972).

57. M. Zatloukal, “Measurements and modeling of temperature-strain rate dependent

uniaxial and planar extensional viscosities for branched LDPE polymer melt,” Polymer

104, 258 (2016).

58. M. Zatloukal, in Annual Technical Conference - ANTEC, Conference Proceedings 3,

2275 (Society of Plastics Engineers, 2013).

59. C. Dae Han, in Rheology in polymer processing 366 (Academic Press, New York, 1976).

60. C. Dae Han, in Rheological Measurement (eds. Collyer, A. A. & Clegg, D. W.) 190

(Chapman & Hall, London, 1998).

61. J. Musil, and M. Zatloukal, “Experimental investigation of flow induced molecular

weight fractionation phenomenon for two linear HDPE polymer melts having same Mn

and Mw but different Mz and Mz+1 average molecular weights,” Chem. Eng. Sci. 81,

146 (2012).

62. R. Pivokonsky, M. Zatloukal, and P. Filip, “On the predictive/fitting capabilities of the

advanced differential constitutive equations for linear polyethylene melts,” J.

Nonnewton. Fluid Mech. 150 (1), 56 (2008).

63. R. Pivokonsky, M. Zatloukal, and P. Filip, “On the predictive/fitting capabilities of the

advanced differential constitutive equations for branched LDPE melts,” J. Nonnewton.

Fluid Mech. 135 (1), 58 (2006).

64. M. Zatloukal, “Differential viscoelastic constitutive equations for polymer melts in

steady shear and elongational flows,” J. Nonnewton. Fluid Mech. 113 (2–3), 209 (2003).

65. C. D. McGrady, C. W. Seay, and D. G. Baird, “Effect of sparse long-chain branching on

the film-casting behavior for a series of well-defined HDPEs,” Int. Polym. Process 24

(5), 428 (2009).

66. K. Aniunoh, and G. M. Harrison, “The processing of polypropylene cast films. I. Impact

of material properties and processing conditions on film formation,” Polym. Eng. Sci. 50

(6), 1151 (2010).

67. G. Lamberti, “Flow-induced crystallization during isotactic polypropylene film casting,”

Polym. Eng. Sci. 51 (5), 851 (2011).

68. M. N. Uvieghara, “The Effect of Deborah Number and Aspect Ratio on the Film Casting

of LLDPE Melts,” Ph.D. thesis, University of Maine, 2004.

69. D. Acierno, L. D. Di Maio, and G. Cuccurullo, “Analysis of temperature fields in film

casting,” J. Polym. Eng. 19 (2), 75 (1999).

70. K. Canning, B. Bian, and A. Co, “Film Casting of a Low Density Polyethylene Melt,” J.

Reinf. Plast. Compos. 20 (5), 366 (2001).

71. G. Lamberti, and G. Titomanlio, “Evidences of flow induced crystallization during

characterized film casting experiments,” Macromol. Symp. 185 (1), 167 (2002).

72. N. Toft, and M. Rigdahl, “Extrusion coating with metallocene-catalysed polyethylenes,”

Int. Polym. Process. 17 (3), 244 (2002).

73. K. K. Aniunoh, and G. M. Harrison, “Experimental Investigation of Film Formation:

Film Casting,” J. Plast. Film Sheeting 22 (3), 177 (2006).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

52

74. S. Kouda, “Prediction of processability at extrusion coating for low-density

polyethylene,” Polym. Eng. Sci. 48 (6), 1094 (2008).

75. V. R. Iyengar, and A. Co, “Film casting of a modified Giesekus fluid: Stability analysis,”

Chem. Eng. Sci. 51 (9), 1417 (1996).

76. M. E. Pis-Lopez, and A. Co, “Multilayer film casting of modified Giesekus fluids Part

1. Steady-state analysis,” J. Nonnewton. Fluid Mech. 66 (1), 71 (1996).

77. S. Smith, and D. F. E. Stolle, “Draw resonance in film casting as a response problem

using a material description of motion,” J. Plast. Film Sheeting 16 (2), 95 (2000).

78. S. Smith, and D. Stolle, “A comparison of Eulerian and updated Lagrangian finite

element algorithms for simulating film casting,” Finite Elem. Anal. Des. 38 (5), 401

(2002).

79. M. Bechert, “Non-Newtonian effects on draw resonance in film casting,” J. Nonnewton.

Fluid Mech. 279, 104262 (2020).

80. Y. L. Yeow, “On the stability of extending films: a model for the film casting process,”

J. Fluid Mech. 66 (3), 613 (1974).

81. G. R. Aird, and Y. L. Yeow, “Stability of film casting of power-law liquids,” Ind. Eng.

Chem. Fundam. 22 (1), 7 (1983).

82. W. Minoshima, and J. L. White, “Stability of continuous film extrusion processes,” J.

Polym. Eng. 2 (3), 211 (1983).

83. N. R. Anturkar, and A. Co, “Draw resonance in film casting of viscoelastic fluids: A

linear stability analysis,” J. Nonnewton. Fluid Mech. 28 (3), 287 (1988).

84. V. R. Iyengar, and A. Co, “Film casting of a modified Giesekus fluid: a steady-state

analysis,” J. Nonnewton. Fluid Mech. 48 (1–2), 1 (1993).

85. P. Barq, J. M. Haudin, J. F. Agassant, and P. Bourgin, “Stationary and dynamic analysis

of film casting process,” Int. Polym. Process. 9 (4), 350 (1994).

86. J. F. Agassant, P. Avenas, J. P. Sergent, and P. J. Carreau, “Polymer Processing:

Principles and Modeling,” (Hanser Gardner Publications, 1991).

87. P. Duffo, B. Monasse, and J. M. Haudin, “Cast film extrusion of polypropylene.

Thermomechanical and physical aspects,” J. Polym. Eng. 10 (1–3), 151 (1991).

88. P. Barq, J. M. Haudin, and J. F. Agassant, “Isothermal and anisothermal models for cast

film extrusion,” Int. Polym. Process. 7 (4), 334 (1992).

89. M. Beaulne, and E. Mitsoulis, “Numerical Simulation of the Film Casting Process,” Int.

Polym. Process. 14 (3), 261 (1999).

90. D. Acierno, L. Di Maio, and C. C. Ammirati, “Film casting of polyethylene terephthalate:

Experiments and model comparisons,” Polym. Eng. Sci. 40 (1), 108 (2000).

91. G. Lamberti, G. Titomanlio, and V. Brucato, “Measurement and modelling of the film

casting process 1. Width distribution along draw direction,” Chem. Eng. Sci. 56 (20),

5749 (2001).

92. G. Lamberti, V. Brucato, and G. Titomanlio, “Orientation and crystallinity in film

casting of polypropylene,” J. Appl. Polym. Sci. 84 (11), 1981 (2002).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

53

93. G. Lamberti, G. Titomanlio, and V. Brucato, “Measurement and modelling of the film

casting process 2. Temperature distribution along draw direction,” Chem. Eng. Sci. 57

(11), 1993 (2002).

94. G. Lamberti, and G. Titomanlio, “Analysis of film casting process: The heat transfer

phenomena,” Chem. Eng. Process. Process Intensif. 44 (10), 1117 (2005).

95. G. Lamberti, and G. Titomanlio, “Analysis of Film Casting Process : Effect of Cooling

during the Path in Air,” Ind. Eng. Chem. Res. 45 (2), 719 (2006).

96. Y.-G. Zhou, W.-B. Wu, J. Zou, and L.-S. Turng, “Dual-scale modeling and simulation

of film casting of isotactic polypropylene,” J. Plast. Film Sheeting 32 (3), 239 (2015).

97. M. Bechert, D. W. Schubert, and B. Scheid, “On the stabilizing effects of neck-in,

gravity, and inertia in Newtonian film casting,” Phys. Fluids 28 (2), 024109 (2016).

98. H. V. Pol, and S. S. Thete, “Necking in Extrusion Film Casting: Numerical Predictions

of the Maxwell Model and Comparison with Experiments,” J. Macromol. Sci. Part B 55

(10), 984 (2016).

99. G. Barot, and I. J. Rao, “Modeling the film casting process using a continuum model for

crystallization in polymers,” Int. J. Non. Linear. Mech. 40 (7), 939 (2005).

100. P. Duffo, B. Monasse, and J.-M. Haudin, “Influence of Stretching and Cooling

Conditions in Cast Film Extrusion of PP Films,” Int. Polym. Process. 5, 272 (1990).

101. D. Silagy, Y. Demay, and J. F. Agassant, “Numerical simulation of the film casting

process,” Int. J. Numer. Methods Fluids 30 (1), 1 (1999).

102. S. Smith, and D. Stolle, “Nonisothermal two-dimensional film casting of a viscous

polymer,” Polym. Eng. Sci. 40 (8), 1870 (2000).

103. S. Smith, and D. Stolle, “Numerical Simulation of Film Casting Using an Updated

Lagrangian Finite Element Algorithm,” Polym. Eng. Sci. 43 (5), 1105 (2003).

104. C. Sollogoub, Y. Demay, and J. F. Agassant, “Cast film problem: A non isothermal

investigation,” Int. Polym. Process. 18 (1), 80 (2003).

105. T. Kajiwara, M. Yamamura, and T. Asahina, “Relationship between Neck-in Phenomena

and Rheological Properties in Film Casting,” Nihon Reoroji Gakkaishi 34 (2), 97 (2006).

106. C. Sollogoub, Y. Demay, and J. F. Agassant, “Non-isothermal viscoelastic numerical

model of the cast-film process,” J. Nonnewton. Fluid Mech. 138 (2–3), 76 (2006).

107. K. Aniunoh, “An Experimental and Numerical Study of the Film Casting Process,” Ph.D.

thesis, Clemson University, 2007.

108. J. S. Lee, and J. M. Kim, “Effect of aspect ratio and fluid elasticity on chain orientation

in isothermal film casting process,” Korean J. Chem. Eng. 27 (2), 409 (2010).

109. S. Shiromoto, Y. Masutani, M. Tsutsubuchi, Y. Togawa, and T. Kajiwara, “A neck-in

model in extrusion lamination process,” Polym. Eng. Sci. 50 (1), 22 (2010).

110. Y. Mu, L. Hang, G. Zhao, X. Wang, Y. Zhou, and Z. Cheng, “Modeling and simulation

for the investigation of polymer film casting process using finite element method,” Math.

Comput. Simul. 169, 88 (2020).

111. H. Kometani, T. Matsumura, T. Suga, and T. Kanai, “Experimental and theoretical

analyses of film casting process,” J. Polym. Eng. 27 (1), 1 (2007).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

54

112. S. D’Halewyu, J. F. Agassant, and Y. Demay, “Numerical simulation of the cast film

process,” Polym. Eng. Sci. 30 (6), 335 (1990).

113. B. Debbaut, J. M. Marchal, and M. J. Crochet, “Viscoelastic effects in film casting,”

Zeitschrift für Angew. Math. und Phys. 46 (SPEC. ISSUE), 679 (1995).

114. K. Sakaki, R. Katsumoto, T. Kajiwara, and K. Funatsu, “Three-dimensional flow

simulation of a film-casting process,” Polym. Eng. Sci. 36 (13), 1821 (1996).

115. H. Zheng, W. Yu, C. Zhou, and H. Zhang, “Three-Dimensional Simulation of the Non-

Isothermal Cast Film Process of Polymer Melts,” J. Polym. Res. 13 (6), 433 (2006).

116. H. Zheng, W. Yu, C. Zhou, and H. Zhang, “Three dimensional simulation of viscoelastic

polymer melts flow in a cast film process,” Fibers Polym. 8 (1), 50 (2007).

117. A. I. . de Waele, “Viscometry and plastometry,” Oil Color Chem. Assoc. J. 6, 33 (1923).

118. W. Ostwald, “Ueber die Geschwindigkeitsfunktion der Viskosität disperser Systeme. I,”

Kolloid-Zeitschrift 36 (2), 99 (1925).

119. R. B. Bird, R. C. Armstrong, and O. Hassager, “Dynamics of Polymeric Liquids, Volume

1: Fluid Mechanics,” (John Wiley, New York, 1987).

120. M. M. Cross, “Rheology of non-Newtonian fluids: A new flow equation for

pseudoplastic systems,” J. Colloid Sci. 20 (5), 417 (1965).

121. P. J. Carreau, “Rheological Equations from Molecular Network Theories,” Trans. Soc.

Rheol. 16 (1), 99 (1972).

122. K. Yasuda, R. C. Armstrong, and R. E. Cohen, “Shear flow properties of concentrated

solutions of linear and star branched polystyrenes,” Rheol. Acta 20 (2), 163 (1981).

123. H. Giesekus, “Die Elastizität von Flüssigkeiten,” Rheol. Acta 5 (1), 29 (1966).

124. H. Giesekus, “A simple constitutive equation for polymer fluids based on the concept of

deformation-dependent tensorial mobility,” J. Nonnewton. Fluid Mech. 11 (1–2), 69

(1982).

125. R. G. Larson, “Constitutive equations for polymer melts and solutions,” (Butterworth-

Heinemann, Boston, 1988).

126. J. M. Wiest, “A differential constitutive equation for polymer melts,” Rheol. Acta 28 (1),

4 (1989).

127. D. Auhl, D. M. Hoyle, D. Hassell, T. D. Lord, O. G. Harlen, M. R. Mackley, and T. C.

B. McLeish, “Cross-slot extensional rheometry and the steady-state extensional response

of long chain branched polymer melts,” J. Rheol. 55 (4), 875 (2011).

128. W. M. H. Verbeeten, G. W. M. Peters, and F. P. T. Baaijens, “Differential constitutive

equations for polymer melts: The extended Pom–Pom model,” J. Rheol. 45 (4), 823

(2001).

129. T. C. B. McLeish, and R. G. Larson, “Molecular constitutive equations for a class of

branched polymers: The pom-pom polymer,” J. Rheol. 42 (1), 81 (1998).

130. A. E. Likhtman, and R. S. Graham, “Simple constitutive equation for linear polymer

melts derived from molecular theory: Rolie–Poly equation,” J. Nonnewton. Fluid Mech.

114 (1), 1 (2003).

131. G. A. J. Holroyd, S. J. Martin, and R. S. Graham, “Analytic solutions of the Rolie Poly

model in time-dependent shear,” J. Rheol. 61 (5), 859 (2017).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

55

132. A. I. Leonov, “Nonequilibrium thermodynamics and rheology of viscoelastic polymer

media,” Rheol. Acta 15 (2), 85 (1976).

133. A. I. Leonov, E. H. Lipkina, E. D. Paskhin, and A. N. Prokunin, “Theoretical and

experimental investigation of shearing in elastic polymer liquids,” Rheol. Acta 15 (7),

411 (1976).

134. A. I. Leonov, and A. N. Prokunin, “An improved simple version of a nonlinear theory of

elasto-viscous polymer media,” Rheol. Acta 19 (4), 393 (1980).

135. A. I. Leonov, and A. N. Prokunin, “On nonlinear effects in the extensional flow of

polymeric liquids,” Rheol. Acta 22 (2), 137 (1983).

136. M. Simhambhatla, and A. I. Leonov, “On the rheological modeling of viscoelastic

polymer liquids with stable constitutive equations,” Rheol. Acta 34 (3), 259 (1995).

137. A. I. Leonov, “Constitutive equations for viscoelastic liquids: Formulation, analysis and

comparison with data,” Rheol. Ser. 8 (C), 519 (1999).

138. R. Pivokonsky, M. Zatloukal, P. Filip, and C. Tzoganakis, “Rheological characterization

and modeling of linear and branched metallocene polypropylenes prepared by reactive

processing,” J. Nonnewton. Fluid Mech. 156 (1–2), 1 (2009).

139. N. Phan-Thien, and R. I. Tanner, “A new constitutive equation derived from network

theory,” J. Nonnewton. Fluid Mech. 2 (4), 353 (1977).

140. N. Phan-Thien, “A Nonlinear Network Viscoelastic Model,” J. Rheol. 22 (3), 259 (1978).

141. R. G. Larson, “A Constitutive Equation for Polymer Melts Based on Partially Extending

Strand Convection,” J. Rheol. 28 (5), 545 (1984).

142. S. A. Khan, and R. G. Larson, “Comparison of Simple Constitutive Equations for

Polymer Melts in Shear and Biaxial and Uniaxial Extensions,” J. Rheol. 31 (3), 207

(1987).

143. A. C. Papanastasiou, L. E. Scriven, and C. W. Macosko, “An Integral Constitutive

Equation for Mixed Flows: Viscoelastic Characterization,” J. Rheol. 27 (4), 387 (1983).

144. X. L. Luo, and R. I. Tanner, “Finite element simulation of long and short circular die

extrusion experiments using integral models,” Int. J. Numer. Methods Eng. 25 (1), 9

(1988).

145. E. Mitsoulis, “Numerical simulation of entry flow of the IUPAC-LDPE melt,” J.

Nonnewton. Fluid Mech. 97 (1), 13 (2001).

146. P. Perzyna, in Advances in Applied Mechanics (eds. Chernyi, G. G. et al.) 9, 243

(Elsevier, 1966).

147. M. M. Denn, C. J. S. Petrie, and P. Avenas, “Mechanics of Steady Spinning of a

Viscoelastic Liquid,” AIChE J. 21 (4), 791 (1975).

148. Y. Demay, “Instabilité d’étirage et bifurcation de Hopf,” Ph.D. thesis, Université de

Nice, 1983.

149. V. R. Iyengar, “Film casting of polymer melts,” Ph.D. thesis, University of Maine, 1993.

150. M. E. Pis-Lopez, and A. Co, “Multilayer film casting of modified Giesekus fluids Part

2. Linear stability analysis,” J. Nonnewton. Fluid Mech. 66 (1), 95 (1996).

151. M. M. Denn, “Fibre Spinning,” Computational Analysis of Polymer Processing (Applied

Science Publishers Ltd., 1983).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

56

152. A. N. Beris, and B. Liu, “Time-dependent fiber spinning equations. 1. Analysis of the

mathematical behavior,” J. Nonnewton. Fluid Mech. 26 (3), 341 (1988).

153. B. Liu, and A. N. Beris, “Time-dependent fiber spinning equations. 2. Analysis of the

stability of numerical approximations,” J. Nonnewton. Fluid Mech. 26 (3), 363 (1988).

154. B. M. Devereux, and M. M. Denn, “Frequency response analysis of polymer melt

spinning,” Ind. Eng. Chem. Res. 33 (10), 2384 (1994).

155. D. K. Gagon, and M. M. Denn, “Computer simulation of steady polymer melt spinning,”

Polym. Eng. Sci. 21 (13), 844 (1981).

156. D. Gelder, “The stability of fiber drawing processes,” Ind. Eng. Chem. Fundam. 10 (3),

534 (1971).

157. R. J. Fisher, and M. M. Denn, “Finite-amplitude stability and draw resonance in

isothermal melt spinning,” Chem. Eng. Sci. 30 (9), 1129 (1975).

158. R. J. Fisher, and M. M. Denn, “A theory of isothermal melt spinning and draw

resonance,” AIChE J. 22 (2), 236 (1976).

159. R. E. Christensen, “Extrusion coating of polypropylene,” SPE J. 18, 751 (1962).

160. J. C. Miller, “Swelling behavior in extrusion,” SPE Trans. 3 (2), 134 (1963).

161. J. P. Sergent, “Etude de deux procédés de fabrication de films. Le soufflage de gaine.

L’extrusion de film à plat.,” Ph.D. thesis, Universite Louis Pasteur, 1977.

162. O. S. Narayanaswamy, “A one-dimensional model of stretching float glass,” J. Am.

Ceram. Soc. 60 (1–2), 1 (1977).

163. G. Lamberti, F. De Santis, V. Brucato, and G. Titomanlio, “Modeling the interactions

between light and crystallizing polymer during fast cooling,” Appl. Phys. A 78 (6), 895

(2004).

164. G. Titomanlio, and G. Lamberti, “Modeling flow induced crystallization in film casting

of polypropylene,” Rheol. Acta 43 (2), 146 (2004).

165. S. Kase, “Studies on melt spinning. IV. On the stability of melt spinning,” J. Appl.

Polym. Sci. 18 (11), 3279 (1974).

166. J. F. Agassant, Y. Demay, C. Sollogoub, and D. Silagy, in International Polymer

Processing 20 (2), 136 (2005).

167. A. Co, in Polymer Processing Instabilities: Control and Understanding (eds.

Hatzikiriakos, S. G. & Migler, K. B.) 287 (CRC Press, 2005).

168. B. Seyfzadeh, G. M. Harrison, and C. D. Carlson, “Experimental studies on the

development of a cast film,” Polym. Eng. Sci. 45 (4), 443 (2005).

169. K. Aniunoh, and G. Harrison, in Annual Technical Conference - ANTEC, Conference

Proceedings 3, 1582 (2007).

170. T. Barborik, and M. Zatloukal, in AIP Conference Proceedings 1843, 030010 (American

Institute of Physics Inc., Zlin, Czech Republic July 26-27, 2017., 2017).

171. C. W. Seay, C. D. McGrady, and D. G. Baird, in Annual Technical Conference - ANTEC,

Conference Proceedings 1380 (2008).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

57

6 TABLES

Table 1. Research work devoted to the experimental investigation of the neck-in phenomenon.

YEAR REFERENCE MATERIAL NOTE

1974 Kase [165] PP Experimental investigation of EFC aimed on process stability.

1986 Dobroth and Erwin [55] LDPE Enlightens the physical background of Edge-beads formation.

1989 Cotto et al. [18] PP Experimental and theoretical investigation of crystalline phase development during

EFC and modelling.

1990 Barq et al. [22] PET The work is aimed on transient phenomena of draw resonance.

1990 Duffo et al. [100] PP Effect of roll temperature on crystallization in EFC process.

1991 Duffo et al. [87] PP Extended experimental and theoretical investigation of crystalline phase

development during EFC and modelling.

1992 Barq et al. [88] PET Experimental and model results are compared, and influence of temperature is

discussed.

1999 Acierno et al. [69] PET; PP

Purely experimental study aimed on film temperature profiles and viscosity.

2000 Acierno et al. [90] PET Role of temperature profile on NI, minor importance of temperature if X < 1/10L0.

2000 Canning and Co [48] LDPE; LLDPE

Purely experimental work deals with effect of rheology, DR and MFR on NI and EB.

2001 Canning et al. [70] LDPE Experimental work capturing velocity, width and thickness profile during EFC.

2001 Lamberti et al. [91] iPP Effect of processing cond. on film development including crystallization.

2001 Satoh et al. [39] LDPE Investigation of viscoelastic effects on NI and edge-beading, relates NI with SH in

uni/pla extensional rate.

2002 Lamberti and Titomanlio [71] iPP Experimental investigation of EFC including width, velocity, temperature and

crystallinity profiles.

2002 Lamberti et al. [92] iPP Experimental measuring and modelling of Hermans orientation factor and

crystallinity. Evaluation of Ziabicki crystallization kinetics from measured film velocity, width and temperature profiles.

2002 Lamberti et al. [93] iPP Experimental measuring of film temperature profiles.

2002 Toft and Rigdahl [72] LDPE; LLDPE;

mLLDPE Experimentally investigates the relationship between polymer elasticity and NI.

2003 Ito et al. [52] LLDPE Experimental oriented work using particle tracking. Confirmation of planar-uniaxial

flows.

2003 Ito et al. [53] LDPE; HDPE; LLDPE

Relates the NI extent to ratio of planar viscosities; axial to transverse.

58

2004 Uvieghara [68] LLDPE The experiment-oriented work, in which the effects of the Deborah number and

the aspect ratio on the EFC were investigated.

2005a Agassant et al. [166] - Film casting review.

2005a Co [167] - Film casting review aimed on the draw resonance.

2005 Lamberti and Titomanlio [94] iPP Experimental part is accompanied by new cooling model with radiant heating.

2006 Aniunoh and Harrison [73] PP Effects of DR and die temperature on temperature, velocity and width profiles.

2006 Bourrigaud et al. [29] LDPE Effect of processing cond. on film development in coating process:

divides De-Dr plane to attainable and unattainable regions.

2006 Lamberti and Titomanlio [95] iPP Effect of processing cond. on film development; film solidification within air gap.

2007 Aniunoh [107] PP Experimentally aimed study on how material properties and process conditions

affect EFC.

2007 Kometani et al. [111] PP;

LDPE Study aims on NI and EB investigation, utilized Giesekus equation was found to be

the most suitable model describing the experimental data.

2007 Shin et al. [36] LDPE; HDPE

Effects of temperature and extensional-thinning and -thickening.

2008 Kouda [74] LDPE Extrusion coating; linking neck-in degree with draw-down force.

2009 Seay and Baird [30] LDPE; LLDPE;

mLLDPE Investigation of effects of LCB and MWD on NI via pom-pom model.

2009 McGrady et al. [65] HDPE; LDPE

Effects of LCB and MWD on NI.

2010 Aniunoh and Harrison [66] PP Effects of Mw, DR and temperature on film formation

2010 Shiromoto et al. [109] LDPE Deals with extrusion lamination process.

Relates the NI gauge to ηE,P/ηE,U.

2010 Shiromoto et al. [54] LDPE Relates the NI gauge to ηE,P/ηE,U.

2011 Lamberti [67] iPP Experimental Study designed to check capabilities of proposed FIC model.

2013 Pol et al. [41] LDPE; HDPE; LLDPE

Effects of LCB and MWD on NI.

2014a Demay and Agassant [23] - Review targeting mainly transient instabilities during EFC.

2014 Pol et al. [42] LDPE; HDPE; LLDPE

Effects of LCB on NI.

2014 Shiromoto [40] LDPE Effect of viscoelastisity on NI, relates the NI gauge to ηE,P/ηE,U.

2015 Chikhalikar et al. [43] PP Effects of LCB on NI.

2015 Zhou et al. [96] iPP Effect of DR on crystallization and development of crystal morphology.

2016 Pol and Thete [98] LDPE; LLDPE

Investigation of NI dependence on De and DR.

2020 Mu et al. [110] PP Influence of processing conditions on film geometry.

59

a) The state of art in EFC reviewing article

Table 2. Articles devoted to modelling the extrusion film casting process: 1D kinematic models. Unless otherwise stated, studies neglect the effects

of flow induced crystallization, inertia, gravity, surface tension, aerodynamic drag, die swell, and film sag.

YEAR REFERENCE MATERIAL EXPERIMENT CONSTITUTIVE

EQATION TRANSIENT

SIMULATION

STRESS BOUNDARY CONDTION

NUMERICAL METHOD

TEMPERATURE NOTE

1974 Yeow [80] Newtonian ✓

Direct numerical

scheme - 4th order Runge-

Kutta

First efforts of EFC modelling from

its stability viewpoint.

1983 Aird and Yeow [81]

Generalized Newtonian

(Power-law) ✓ ?

Aims on investigation of stability EFC process.

1983 Minoshima and

White [82] Newtonian ✓ ? ✓

Investigation of stability of EFC process by model incorporating

thermal effects.

1988 Anturkar and Co

[83]

Generalized UCM with deformation rate dependent η

and λ

✓ 3 ?

Stemming from (Yeow 1974) work, investigation of effects of viscoelasticity on the draw resonance phenomenon.

1990 Barq et al. [22] PET ✓ Newtonian ✓ FDM The work is aimed on transient phenomena of draw resonance.

1991a Alaie and

Papanastasiou [45] compared

to PP

K-BKZ model with PSM damping

function 0

FEM Galerkin

Effects of die design and melt rheology, including temperature on

film thickness in EFC process.

1993 Iyengar and Co [84]

mod. Giesekus with finite chain

stretch 4

4th order Runge-Kutta + adaptive step size control

Follows (Anturkar and Co 1988) work. Theoretical investigation of extensional viscosity curve and its effects on film velocity and stress

profiles.

1994 Barq et al. [85] compared

to PET UCM ✓ 3 FDM

Theoretical investigation of limits of EFC process regarding to DRc.

1996 Iyengar and Co [75]

mod. Giesekus with finite chain

stretch ✓ 4

4th order Runge-Kutta + adaptive step size control

Utilization of linear stability analysis to determination of DRc,

examination of effects extensional thickening and thinning.

60

1996 Pis-Lopez and Co

[76]

mod. Giesekus with finite chain

stretch 4

4th order Runge-Kutta + adaptive step size control

Steady-state analysis, multilayer film casting, study on the effect rheological properties and

processing conditions on film velocity and stress profiles.

2000 S. Smith and Stolle

[77]

Hooke’s Law with creep

✓ Lagrangian

FEM

Except few steady-state results, the work is focused on DRc

determination viewed as a response problem.

2002 S. Smith and Stolle

[78]

Hooke’s Law with creep

✓ 0 FEM + Newton

Raphson

Comparison of Eulerian and Updated Lagrangian FE algorithms for film

casting simulation.

2015 Polychronopoulos and Papathanasiou

[26]

Newtonian; Linear PTT

0 OpenFOAM -

FVM

Effect of draw ratio on die swell in film casting.

2020 Bechert [79]

Giesekus; exponential PTT;

Generalized Newtonian

(Carreau-Yasuda);

✓ 4 AUTO-07P, MATLAB - CHEBFUN

Investigation of influence of viscoelastic, and in particular

non-Newtonian, effects on the draw resonance instability in EFC.

a) Discrete relaxation spectra were used in this study.

Legend to the stress boundary condition column on the die:

0 - all stress components are set to zero considering entire stress relaxation due to the die swell phenomenon.

3 - the die exit stress state is given by the axial upstream extra stress component.

4 - the thickness to axial extra stress component ratio for the upstream/down-stream side.

61

Table 3. Articles devoted to modelling the extrusion film casting process: 1.5D kinematic models. Unless otherwise stated, studies neglect the

effects of flow induced crystallization, inertia, gravity, surface tension, aerodynamic drag, die swell, and film sag.

YEAR REFERENCE MATERIAL EXPERIMENT CONSTITUTIVE

EQATION MULTI-MODE

APPROACH

STRESS BOUNDARY CONDTION

NUMERICAL METHOD

TEMPERATURE CRYSTALLIZATION NOTE

1989a Cotto et al. [18] PP ✓ Newtonian FDM ✓ ✓

Experimental and theoretical investigation of crystalline phase

development during EFC and modelling.

1990 a Duffo et al. [100] PP ✓ Newtonian FDM ✓ ✓ Effect of roll temperature on crystallization in EFC process.

1991a Agassant et al.

[86] Newtonian FDM

Study deals with derivation of film width-variation model.

1991a Duffo et al. [87] PP ✓ Newtonian FDM ✓ ✓

Extended experimental and theoretical investigation of

crystalline phase development during EFC and modelling.

1992ac Barq et al. [88] PET ✓ Newtonian

Runge-Kutta and

Adams-Bashforth's

method

✓ Experimental and model results are

compared, and influence of temperature is discussed.

1996b Silagy et al. [27] Newtonian;

UCM 1

4nd order Runge-Kutta; linear stability

analysis

Brings afterward widely utilized 1.5D model based on (Narayanaswamy

1977) kinematics. Study of unattainable zone.

1999 Beaulne and

Mitsoulis [89]

compared to PP; PET;

LDPE;

Newtonian; UCM; K-BKZ (PSM) model

✓ 1 4nd order

Runge-Kutta; FEM;

✓ Simulation result comparison with

many related studies.

1999 Kwon [25] Typical

values for PE

Leonov ✓ 5 ?

Qualitative simulation of anisotropic die swelling behavior of the extruded

film

2000a Acierno et al.

[90] PET ✓ Newtonian

Ordinary single step

Eulerian algorithm

Role of temperature profile on NI; minor importance of temperature if

X < 1/10L0.

2001 Lamberti et al.

[91] iPP ✓

Generalized Newtonian

(Cross)

Euler’s method

✓ ✓

Effect of processing cond. on film development including

crystallization.

62

2002 Lamberti et al.

[93] iPP ✓

Generalized Newtonian

(Cross) FDM ✓ ✓

Experimental measuring of film temperature profiles.

2003 Ito et al. [53] LDPE; HDPE; LLDPE

✓ Newtonian;

UCM

Shooting method

Relates the NI extent to ratio of

planar viscosities; axial to transverse.

2005 Lamberti and

Titomanlio [94] iPP ✓ UCM ✓

Experimental part is accompanied by new cooling model with radiant

heating.

2005 Barot and Rao

[99] compared

to iPP UCM

Variable order method

✓ ✓ Modelling of crystallization during

EFC process.

2006 Bourrigaud et al.

[29] LDPE ✓ UCM 0

Num. integration+

Newton method

Effect of processing cond. on film development in coating process: divides De-DR plane to attainable

and unattainable regions.

2013 Pol et al. [41] LDPE; HDPE; LLDPE

✓ XPP; RP-S

✓ 2 MATLAB -

ode15s Effects of LCB and MWD on NI.

2014 Pol et al. [42] LDPE; HDPE; LLDPE

✓ XPP; RP-S

✓ 2 MATLAB -

ode15s ✓ Effects of LCB on NI.

2015 Chikhalikar et al.

[43] PP ✓

XPP; RP-S

✓ 1 MATLAB -

ode15s ✓ Effects of LCB on NI.

2015 Zhou et al. [96] iPP ✓

Generalized Newtonian

(Cross)

FDM/Monte Carlo

✓ ✓ Effect of DR on crystallization and

development of crystal morphology.

2016bc Bechert et al.

[97] Newtonian AUTO-07P

Study reveals a stabilizing effect of NI, gravity and inertia on EFC process

from draw resonance viewpoint.

2016 Pol and Thete

[98] LDPE; LLDPE

✓ UCM 0

Finite difference

scheme; time marching scheme

Investigation of NI dependence on

De and DR.

2017 Barborik et al.

[37] compared

to LDPE mod. Leonov 5

4th order Runge-Kutta

Investigation of the role of planar to uniaxial extensional viscosity ratio, extensional strain hardening and

Deborah number.

2017b Dhadwal et al.

[47] exponential PTT ✓ 1

MATLAB - ode15s;

linear stability analysis

Effects of melt relaxation modes on

process stability.

2017 Thete et al. [44] LDPE; LLDPE

✓ UCM;

exponential PTT ✓ 2

MATLAB - ode15s

Effects of melt relaxation modes on

NI.

63

2018 Barborik and Zatloukal [38]

compared to LDPE

mod. Leonov 5 4th order

Runge-Kutta

Study aimed on the effect of the second to first normal stress difference at the die exit and

Deborah number.

2019 Barborik and Zatloukal [35]

compared to iPP

mod. Leonov 5 4th order

Runge-Kutta ✓ ✓

Theoretical investigation of crystalline phase development

during EFC.

a) The model prediction capabilities are limited only to the final value of the film width L(X) to be in the range of ( )0 0 2 −L L X L X .

b) The study also includes a transient simulation.

c) The effects of inertia and gravity are included in the model.

Legend to the stress boundary condition column on the die:

0 - all stress components are set to zero considering entire stress relaxation due to the die swell phenomenon.

1 - at least one stress component is given by the Newtonian solution for the downstream side, independent of the type of constitutive equation

used.

2 - two extra stress components are set manually without further justification.

5 - ratio of the second to the first normal stress difference, −N2/N1, calculated from the upstream side by using a viscoelastic constitutive equation.

64

Table 4. Articles devoted to modelling the extrusion film casting process: 2D kinematic models. Unless otherwise stated, studies neglect the effects

of surface tension, aerodynamic drag, die swell, and film sag.

YEAR REFERENCE MATERIAL EXPERIMENT MODEL

DIMENSIONALITY CONSTITUTIVE

EQATION MULTI-MODE

APPROACH TRANSIENT

SIMULATION

STRESS BOUNDARY CONDTION

NUMERICAL METHOD

TEMPERATURE CRYSTALLIZATION FLOW-INDUCED

CRYSTALLIZATION INERTIA GRAVITY NOTE

1990 D’Halewyu et

al. [112] 2D Newtonian FEM; FVM

First 2D numerical model

of EFC.

1995 Debbaut et al.

[113] 2D

Newtonian; Generalized Newtonian

(Power-law); UCM;

Giesekus

0 FEM

Examining edge-beading in

the light of viscoelastic flow

behavior.

1997 W. S. Smith

[12] compared

to PP

1D & 2D

Newtonian Eulerian

FEM ✓ ✓ ✓

Aims on effects of inertia, gravity and temperature

on film formation;

model comparison.

1998 Silagy et al.

[28]

compared to LLDPE;

LDPE

1.5D & 2D

Newtonian; UCM

✓ 0, 1

2nd order Runge-

Kutta; FEM; linear

stability analysis for

2D

Effect of DR, A and De on NI, comparison of models results; investigation of

unattainable region.

1999 Silagy et al.

[101] compared

to PES

1D & 2D

Newtonian ✓

FEM – Galerkin method;

linear stability

analysis for 1D and 2D

Effect of DR and A on NI, and

onset of draw resonance.

2000 Christodoulou

et al. [46] PET 2D

exponential PTT

✓ ✓ multi FEM EVSS

Effect of stress boundary

conditions on computational

stability; determination of

critical DR.

65

2000 S. Smith and Stolle [102]

Typical values for

PET; LDPE

2D Newtonian FEM ✓ ✓ ✓

Investigation of factors that

contribute NI reduction (self-

weight, EB, cooling).

2001 Satoh et al.

[39] LDPE ✓ 2D

Newtonian; Larson

✓ 1 Galerkin

FEM ✓

Investigation of viscoelastic

effects on NI and edge-beading, relates NI with SH in uni/pla

extensional rate.

2001 W. S. Smith

[8]

Typical values for

PP

1.D; 1.5D &

2D

Generalized Newtonian

(Power-Law); UCM

✓ 0 UL FEM ✓

Model comparison –

Eularian vs Lagrangian approach.

2003 S. Smith and Stolle [103]

Typical values for

PP 2D

Hooke’s Law with creep

✓ 0 UL FEM ✓

Based on Lagrangian

description of motion, DRc

determined from response problem.

2003 Sollogoub et

al. [104] compared

to PET 2D Newtonian

Quadrangle continuous

Galerkin FEM

✓ ✓

Investigation of effects of HTC on

film development.

2006 Kajiwara et al.

[105] 2D

Newtonian; Generalized Newtonian

(Cross); Larson;

exponential PTT

1 Galerkin

FEM ✓

Relates NI to ratio of uniaxial

to planar extensional

viscosity.

2006 Sollogoub et

al. [106] 2D UCM 0

Quadrangle continuous

Galerkin FEM

Effect of HTC and VE on film formation.

2007 K. Aniunoh

[107] PP ✓ 2D Giesekus Matlab ✓ ✓ ✓

Experimentally aimed study on how material

properties and process

66

conditions affect EFC.

2007 Kometani et

al. [111] PP;

LDPE ✓ 2D

Newtonian; Generalized Newtonian

(Bird Carreau); Giesekus

Polyflow

FEM

Study aims on NI and EB

investigation, utilized Giesekus

equation was found to be the most suitable

model describing the experimental

data.

2007 Shin et al.

[36] LDPE; HDPE

✓ 2D exponential

PTT ✓ 1

FEM ALE

Effects of temperature and

extensional-thinning

and -thickening.

2010 Lee and Kim

[108] 2D

exponential PTT

✓ Galerkin

FEM

Investigation of high aspect

ratios causing the highly oriented

molecular structures.

2010 Shiromoto et

al. [109] LDPE ✓ 2D

exponential PTT

ANSYS,

Polyflow FEM

Deals with extrusion

lamination process.

Relates the NI gauge to ηE,P/ηE,U.

2010 Shiromoto et

al. [54] LDPE ✓ 2D

exponential PTT

ANSYS,

Polyflow FEM

Relates the NI gauge to ηE,P/ηE,U.

2014 Shiromoto

[40] LDPE ✓ 2D

Generalized Newtonian (Carreau-Yasuda);

exponential PTT

✓ ANSYS,

Polyflow FEM

Effect of viscoelastisity on NI, relates the NI

gauge to ηE,P/ηE,U.

2020 Mu et al.

[110] PP ✓ 2D

Generalized Newtonian (Carreau)

FEM – with standard Galerkin

formulation

Influence of processing

conditions on film geometry.

67

Legend to the stress boundary condition column on the die:

0 - all stress components are set to zero considering entire stress relaxation due to the die swell phenomenon.

1 - at least one stress component is given by the Newtonian solution for the downstream side, independent of the type of constitutive equation

used.

Table 5. Articles devoted to modelling the extrusion film casting process: 3D kinematic models. Unless otherwise stated, studies neglect the

effects of crystallization, flow induced crystallization, surface tension, aerodynamic drag, die swell, and film sag.

YEAR REFERENCE MATERIAL EXPERIMENT CONSTITUTIVE

EQATION

STRESS BOUNDARY CONDTION

NUMERICAL METHOD TEMPERATURE INERTIA GRAVITY NOTE

1996 Sakaki et al.

[114] Newtonian Streamline FEM Effects of DR, X, L0 on NI and EB.

2006 Zheng et al.

[115] compared to

PET

Generalized Newtonian (Carreau)

FEM ✓ ✓ Includes Shear-thinning, self-weight,

viscous dissipation.

2007 Zheng et al.

[116] compared to

PET exponential PTT 1 Polyflow FEM ✓ ✓

Effects of strain-hardening and elasticity on the final film shape.

Legend to the stress boundary condition column on the die:

1 - at least one stress component is given by the Newtonian solution for the downstream side, independent of the type of constitutive equation

used.

68

Table 6. Influence of material and extrusion film casting process variables on the extent of neck-in phenomenon (opposite trends are highlighted

in bold).

YEAR REFERENCE MATERIAL

DEPENDENT VARIABLE

INDEPENDENT VARIABLE

Neck-in

Strain hardening in uniaxial extension

Mw, MWD

(λ)

Draw ratio (adjusted via

chill roll speed, vx(X))

Draw ratio (adjusted via

melt velocity at the die exit,

vx(0))

vx(X) and vx(0) (draw ratio is

kept constant)

Take-up length

Temperature –N2/N1

(if De is high) ηE,P/ηE,U

1986 Dobroth and Erwin [55] LDPE — — — — — — —

1999 Acierno et al. [69] PP, PET — — — — — —

2000 Acierno et al. [90] PET — — — — — — —

2000 Canning and Co [48] LDPE, LLDPE

— c (LDPE)

(LLDPE) — — — — —

2001 Canning et al. [70] LDPE — — —/ — — — — — —

2001 Satoh et al. [39] LDPE — c — — — — —

2001 Lamberti et al. [91] iPP — — a — — — — —

2002 Lamberti et al. [93] iPP — — — a — — — — —

2002 Lamberti and Titomanlio [71] iPP — — — a — — — — —

2002 Lamberti et al. [92] iPP — — — a — — — — —

2002 Toft and Rigdahl [72] LDPE, LLDPE,

mLLDPE b — — — —

2003 Ito et al. [53] LDPE, LLDPE, HDPE

— — (LLDPE, HDPE)

c (LDPE) — — — — —

2003 Ito et al. [52] mLLDPE — — — — — — — —

2005 Seyfzadeh et al. [168] PET — — — — — — —

2006 Aniunoh and Harrison [73] PP — — — — — —

2006 Lamberti and Titomanlio [95] iPP — — — — — — — —

2007 Aniunoh and Harrison [169] PP — — — — — — — —

2007 Kometani et al. [111] PP,

LDPE — —

(PP)

—/c (LDPE) — — — — — —

2008 Kouda [74] LDPE — — — — — — — —

2009 McGrady et al. [65] HDPE — — — — — — — —

2009 Seay and Baird [30] LDPE, LLDPE,

mLLDPE — — — — — —

2010 Aniunoh and Harrison [66] PP — — — —

69

2010 Shiromoto et al. [109] LDPE — — — — — 2010 Shiromoto et al. [54] LDPE — — — — — —

2013 Pol et al. [41]

LDPE, LLDPE,

mLLDPE, HDPE

— — — — —

2014 Pol et al. [42] LDPE, LLDPE

— — — — — —

2015 Chikhalikar et al. [43] PP — — — — — —

2015 Zhou et al. [96] iPP — — — — — — — —

2018 Barborik and Zatloukal [38] LDPE — — — — — —

2020 Mu et al. [110] PP — — — — — — —

a) The observed opposite trend is due to the crystallization of the polymer in the drawing region.

b) Used LLDPE and mLLDPE (octene/hexene comonomer types) have a high Trouton ratio (>20), indicating the presence of LCB.

c) The effect of the draw ratio on the neck-in is complex for LDPEs (most likely due to the non-monotonic dependence of the uniaxial and

planar extensional viscosities on the extensional strain rate that is typical for branched polymers).

70

7 LIST OF SYMBOLS

Latin Symbols Meaning Unit

A Aspect ratio 1

a Parameter in Cross and Carreau-Yasuda model 1

B Bead ratio 1

b Dissipation term in modified Leonov model s-1

cb Chain extensibility parameter in modified

Giesekus model

1

c Recoverable Finger tensor 1

1C

− Finger strain tensor 1

c

Jaumann (corotational) time derivative of the

recoverable Finger strain tensor in modified

Leonov model

s-1

D Deformation rate tensor s-1

De Deborah number 1

DR Draw ratio 1

CDR Critical draw ratio 1

E Elasticity matrix in Hooke’s Law Pa

e Half-thickness of the film at any x location m

0e Die half-gap (half-thickness of the film at the

die exit)

m

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

71

pe

Irreversible rate of strain tensor in modified

Leonov model

s-1

( )f x Rate of deformation in transverse y-direction s-1

G Linear Hookean elastic modulus Pa

G Storage modulus Pa

G Loss modulus Pa

( )g x Rate of deformation in thickness z-direction s-1

HTC Heat transfer coefficient J·s-1·K-1·m-2

edge

fh Edge final film thickness mm

center

fh Center final film thickness mm

DII Second invariant of deformation rate tensor s-1

cI First invariant of recoverable Finger tensor 1

cII Second invariant of recoverable Finger tensor 1

1C

I − First invariant of the Finger strain tensor 1

1C

II − Second invariant of the Finger strain tensor 1

i Index i, noting the spatial direction 1

0

EJ Linear steady-state elastic compliance Pa-1

j Relaxation mode identification number 1

L Half-width of the film at any x location m

vL Velocity gradient tensor s-1

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

72

0L Half-width of the die (half-width of the film at

the die exit)

m

m Flow consistency index in Power-law model Pa.sn

MFR , m Mass flow rate kg·h-1

Mn Number average molar mass g·mol-1

Mw Mass average molar mass g·mol-1

N Presents the highest value available –

n Flow behavior index in GNM 1

Ln Non-linear Leonov model parameter 1

n

Adjustable parameter in relaxation time

function in Generalized UCM model

1

NI Maximum attainable neck-in m

1N First normal stress difference Pa

2N Second normal stress difference Pa

2 1N N− Stress state at the die exit 1

q Number of arms in the XPP model 1

T Melt temperature °C

t Present time in K-BKZ model s

t Past time in K-BKZ model s

xv

Axial velocity component of the film at any x

location

m·s-1

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

73

( )xv 0 Axial velocity component at the die exit

(velocity in the machine direction)

m·s-1

xv (X) Chill roll speed (Take-up rate) m·s-1

yv Velocity component of the film in transverse

y-direction at any x location

m·s-1

zv Velocity component of the film in thickness

z-direction at any x location

m·s-1

W Elastic potential in modified Leonov model Pa

X Take-up length (drawing distance, air gap) m

x, y, z Spatial coordinates in axial, transverse and

thickness direction, respectively

m

x Position in axial x-direction m

Z Function in modified Giesekus model 1

Greek Symbols Meaning Unit

Anisotropy parameter in Giesekus and XPP

model

1

i Parameter in K-BKZ model 1

Non-linear Leonov model parameter 1

c Convective constraint release coefficient in

RP-S model

1

i Parameter in K-BKZ model 1

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

74

Shear strain rate s-1

Total strain increment in Hooke’s Law 1

c

Modified Perzyna creep strain increment in

Hooke’s Law

1

Stress increment in Hooke’s Law Pa

The Kronecker delta, unit tensor 1

0 Fitting scalar parameter in RP-S model 1

p Parameter in PTT model 1

p Extensional strain rate s-1

, s Steady shear viscosity Pa·s

0 Newtonian viscosity, zero-shear viscosity Pa·s

b Steady biaxial extensional viscosity Pa·s

p Polymer viscosity in Giesekus model Pa·s

Infinite-shear-rate viscosity, solvent viscosity in

Giesekus model

Pa·s

E,P , P Steady planar extensional viscosity Pa·s

E,U , U Steady uniaxial extensional viscosity Pa·s

E,U,max Maximal steady uniaxial extensional viscosity Pa·s

E,U,max

0

η

3η Uniaxial extensional strain hardening 1

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

75

Parameter in K-BKZ model 1

Backbone tube stretch in XPP model 1

, 1 Melt relaxation time s

Average relaxation time s

0 Adjustable parameter in relaxation time

function in Generalized UCM model

s

0b Orientation relaxation time in XPP model s

0s Stretch relaxation time in XPP model s

d Reptation relaxation time in RP-S model s

r Rouse relaxation time in RP-S model s

t Adjustable parameter in relaxation time

function in Generalized UCM model

s

( )1−

relaxation time tensor in XPP model s

Non-linear Leonov model parameter 1

Non-linear Leonov model parameter 1

p Parameter in PTT model 1

Non-linear model parameter in Larson model 1

Extra stress tensor Pa

Gordon–Schowalter convected time derivative

of the stress tensor in PTT model

Pa·s-1

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

76

Upper-convected time derivative of the stress

tensor

Pa·s-1

p

Polymer contribution to the stress tensor in

Giesekus model

Pa

s

Solvent contribution to the stress tensor in

Giesekus model

Pa

xx Normal stress in axial x-direction Pa

xx Dimensionless normal stress in axial x-direction 1

yy Normal stress in transverse y-direction Pa

yy

Dimensionless normal stress in transverse

y-direction

1

zz Normal stress in thickness z-direction Pa

zz

Dimensionless normal stress in thickness

z-direction

1

Gradient operator 1

Latin Abbreviations and Acronyms Unit

1D,1.5D, 2D and 3D Model dimensionality, e.g. 1D –

One-dimensional model

CHEBFUN Framework within the MATLAB –

DE Doi-Edwards integral tube model –

EB Edge-beading phenomenon –

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

77

EFC Extrusion film casting –

FDM Finite difference method –

FE Finite element –

FEM Finite element method –

FEM ALE FEM the arbitrary Lagrangian Eulerian method –

FEM EVSS FEM the elastic viscous stress splitting –

FIC Flow-induced crystallization –

FVM Finite volume method –

GNM Generalized Newtonian model –

HDPE Material: high-density polyethylene –

iPP , PP Material: isotactic polypropylene –

K BKZ−

Kaye-Bernstein-Kearsley-Zapas constitutive

model

LCB Long chain branching –

LDPE Material: low-density polyethylene –

LLDPE Material: linear low-density polyethylene –

MWD Molecular weight distribution –

mLLDPE

Material: linear metallocene-catalyzed

low-density polyethylene

ode15s Differential equation solver within MATLAB

software

PBAT

Material:

poly(3-hydroxybutyrate-co-3-hydroxyvalerate)

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

78

PBS Material: polybutylene succinate –

PES Material: polyethersulfone –

PET Material: polyethylene terephthalate –

PLA Material: polylactide –

PP Material: polypropylene –

PS Material: polystyrene –

PSM Papanastasiou-Scriven-Macosko damping

function for K-BKZ constitutive model

PTT Phan–Thien and Tanner constitutive model –

RP S− Rolie-Poly Strech constitutive model –

SH Strain hardening –

( )tr Denotes the trace of a matrix –

UCM Upper convected Maxwell constitutive model –

UL FEM Updated Lagrangian FEM –

VE Viscoelasticity –

XPP eXtended Pom–Pom constitutive model –

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

79

8 FIGURES

Fig. 1: Schematics of the extrusion film casting kinematics. Reproduced from [T.

Barborik, and M. Zatloukal, in AIP Conference Proceedings 1843, 030010

(American Institute of Physics Inc., Zlin, Czech Republic July 26-27, 2017.,

2017)], [170], with the permission of AIP Publishing.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

80

Fig. 2: Visualization of the effect of draw resonance on film width and thickness.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

81

Fig. 3: Visualization of the neck-in phenomenon during extrusion film casting.

Reproduced with permission from Int. J. Heat Mass Transf. 111, 1296

(2017) [37]. Copyright 2017, Elsevier.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

82

Fig. 4: Effect of Carreau relaxation time due to increased MWD on the neck-in for

linear metallocene-catalyzed low-density polyethylene (mLLDPE, = 0.33

s at 150 °C, Mw = 110 000 g/mol, Mw/Mn = 2.04, LCB/10 000 C=0) and

Zieglar-Natta-catalyzed linear low-density polyethylene LLDPE ( = 1.62

s, Mw = 122 700 g/mol, Mw/Mn = 3.44, LCB/10 000C=0). The process

conditions for both samples were following: die width was 101.6 mm and

its gap size was 0.57 mm, take-up length equal to 141.4 mm, a temperature

of 150 °C, an extrusion shear rate of 8.62 s-1 and a drawdown ratio of 15.

The flow direction is oriented left to right here. Selected and digitalized

from [30].

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

83

Fig. 5: Effect of the longest relaxation time (i.e. 0

0 EJ = , where 0

EJ is the linear steady-

state elastic compliance) due to the increased molecular weight on the neck-in for linear

polypropylenes PP1 ( = 15.5 s at 230 °C, Mw = 527 000 g/mol, Mw/Mn = 5.3) and

PP2 ( = 4.2 s at 230 °C, Mw = 359 000 g/mol, Mw/Mn = 6.9) at two draw down ratios

(10 and 35). The process conditions for both samples were as follows: the velocity at

the die exit was 1.3mm/s, the take-up length was 130 mm, the temperature was 200 °C

[31–34]. Here, the flow direction is oriented from left to right. Reproduced from [H. Münstedt,

in AIP Conference Proceedings (eds. Zatloukal, M. & Musil, J.) 2107, 030001 (American

Institute of Physics Inc., Zlin, Czech Republic, 2019)], [34], with the permission of AIP

Publishing.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

84

Fig. 6: Effect of long chain branching on the neck-in development at different draw

ratios, experimental data: 6a) LDPE with LCB; 6b) LLDPE without LCB.

The process conditions for both samples were following: the die width was

254 mm and the temperature 240 °C. Here the flow direction is oriented

from top to bottom. Reproduced with permission from J. Plast. Film

Sheeting 16 (3), 188 (2000) [48]. Copyright 2000, SAGE Publications.

6a)

6b)

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

85

Fig. 7: Effect of increased long chain branching on the neck-in: 7a) linear

metallocene-catalyzed low-density polyethylene (mLLDPE); 7b) highly

branched low-density polyethylene (LDPE). The process conditions for both

samples were as follows: the die width was 101.6 mm and its gap size 0.57

mm, the take-up length 141.4 mm, the temperature 150 °C, the extrusion

shear rate 1.33 s-1 and the drawdown ratio 10. Here, the flow direction is

oriented from top to bottom. Courtesy of professor Donald G. Baird for his

permission to reprint this figure from [171].

7b)

7a)

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

86

Fig. 8: Visualization of the polymer flow field in the air gap during film casting of

linear metallocene-catalyzed low-density polyethylene (mLLDPE, Mw =

57 200 g/mol, Mw/Mn = 2.26), in which the flow direction is from top to

bottom. The process conditions were as follows: the velocity at the die exit

was 10.7 mm/s, the take-up length equal to 150 mm, the temperature

190 °C. Selected and digitalized from [52].

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

87

Fig. 9: Maximum attainable normalized neck-in (i.e. NI/X) for different LDPEs.

Experimental data and proposed analytical model predictions (Eq. 6) are

presented here with open and filled symbols, respectively. Reproduced with

permission from J. Nonnewton. Fluid Mech. 255, 39 (2018) [38]. Copyright

2018, Elsevier.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

88

Fig. 10: Schematic illustration of the extrusion film casting process with the indicated

film cross-sectional development (formation of edge-beads) in the air-gap.

The curves within the film represent the borders between the planar

extensional flow (center area) and the uniaxial extensional flow (edge

areas).

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

89

Fig. 11: Evolution of edge-beading at different draw ratios, experimental data for

LLDPE without LCB (corresponding film width profiles are provided in

Figure 6b). Process conditions were following: the die width was 254 mm

and the temperature 240 °C. Reproduced with permission from J. Plast.

Film Sheeting 16 (3), 188 (2000) [48]. Copyright 2000, SAGE Publications.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

90

Fig. 12: Bead ratio as a function of DR for LDPE at 177 °C. The symbols represent

experimental data, while the line represents the theoretical value predicted

by the Eq. 7. Reproduced with permission from Polym. Eng. Sci. 26 (7), 462

(1986) [55]. Copyright 1986, John Wiley and Sons.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

91

Fig. 13: Film temperature and width against dimensionless drawing distance for PET

for DR = 10 (A2 experiment) and DR = 20 (A4 experiment), die width =

200 mm and X = 150 mm. The simulations are based on the Newtonian

constitutive equation and 1.5D kinematic model. Reproduced with

permission from Polym. Eng. Sci. 40 (1), 108 (2000) [90]. Copyright 2000,

John Wiley and Sons.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

92

Fig. 14: Effect of elasticity on the edge-beading for DR = 50 predicted by the 2D film

casting model considering Newtonian and UCM constitutive equations. The

vertical axis represents the dimensionless film thickness (actual thickness

divided by the die gap) and the horizontal axis represents the dimensionless

film width (actual distance from the film center divided by the die width).

Reproduced with permission from J. Nonnewton. Fluid Mech. 79 (2–3), 563

(1998) [28]. Copyright 1998, Elsevier.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

93

Fig. 15: Comparison between experimentally measured uniaxial (open symbols) and

planar (closed symbols) extensional viscosities together with the multimode

Pom-Pom model predictions for the uniaxial (solid lines) and planar

(dashed lines) extensional viscosity for the LDPE series. Reproduced from

[D. Auhl, D. M. Hoyle, D. Hassell, T. D. Lord, O. G. Harlen, M. R.

Mackley, and T. C. B. McLeish, “Cross-slot extensional rheometry and the

steady-state extensional response of long chain branched polymer melts,”

J. Rheol. 55 (4), 875 (2011)], [127], with the permission of AIP Publishing.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

94

Fig. 16: Normalized film width vs. draw ratio for branched LDPE at three different

air-gaps (squares: X = 10mm; circles: X = 90mm; inverted triangles:

X = 228mm; die width = 100 mm). The symbols represent experimental

data and the lines are predictions based on the XPP constitutive equation

and the 1.5D kinematic model. Reproduced from [H. V. Pol, S. S. Thete, P.

Doshi, and A. K. Lele, “Necking in extrusion film casting: The role of

macromolecular architecture,” J. Rheol. 57 (2), 559 (2013)], [41], with the

permission of AIP Publishing.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

95

Fig. 17: Normalized film width vs. draw ratio for linear LLDPE at three different air-

gaps (squares: X = 10 mm; circles: X = 90 mm; inverted triangles:

X = 228 mm; die width = 100 mm). The symbols represent experimental

data and the lines are predictions based on the RP-S constitutive equation

and the 1.5D kinematic model. Reproduced from [H. V. Pol, S. S. Thete, P.

Doshi, and A. K. Lele, “Necking in extrusion film casting: The role of

macromolecular architecture,” J. Rheol. 57 (2), 559 (2013)], [41], with the

permission of AIP Publishing.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

96

Fig. 18: Normalized film width vs. drawing distance at DR = 34.7, die

width = 200 mm and X = 400 mm for linear PP (left) and normalized film

width vs. DR at die width = 100 mm at X = 230 mm for branched LDPE

(right). The symbols represent experimental data and the lines are

predictions based on the modified Leonov constitutive equation and the

1.5D kinematic model. Left figure reproduced from [T. Barborik, and M.

Zatloukal, “Effect of Heat Transfer Coefficient, Draw Ratio and Die Exit

Temperature on the Production of Flat Polypropylene Membranes,” Phys.

Fluids 31 (5), 053101 (2019)], [35], with the permission of AIP Publishing.

Right figure reproduced with permission from Int. J. Heat Mass Transf.

111, 1296 (2017) [37]. Copyright 2017, Elsevier.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

97

Fig. 19: The planar to uniaxial extensional viscosity ratio predicted by the PTT model

for three different LDPEs plotted as a function of the extensional strain

rate. Reproduced with permission from Polym. Eng. Sci. 50 (1), 22 (2010)

[109]. Copyright 2010, John Wiley and Sons.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

98

Fig. 20: Comparison of the film edge shapes for three LDPEs with different uniaxial

extensional strain hardening (a-high, b-middle, c-low uniaxial extensional

strain hardening) where die width = 600 mm, and DR was kept the same,

equal to 40 by proper adjusting of vx(0) and vx(X). Left: the role of take-up

velocities (80, 120 and 190 m/min), X = 160 mm; Right: the role of air-gaps

(160, 190 and 220 mm) at the fixed take-up velocity (120 m/min). The

symbols represent experimental data and the lines are predictions based on

the multimode PTT constitutive equation and the 2D kinematic model.

Reproduced with permission from Polym. Eng. Sci. 50 (1), 22 (2010) [109].

Copyright 2010, John Wiley and Sons.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

99

Fig. 21: Film width vs. drawing distance (left) and final film thickness across the film

width (right) at DR = 32.7, die width = 0.25 m and X = 0.09 m for

branched LDPE (right). Open circles represent experimental data and lines

(or lines with filled circles) are predictions based on the Larson constitutive

equation and the 2D kinematic model. Reproduced with permission from

Polym. Eng. Sci. 41 (9), 1564 (2001) [39]. Copyright 2001, John Wiley and

Sons.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

100

Fig. 22: Uniaxial extensional viscosity vs. extensional strain rate predicted by the

multimode PTT constitutive equation for LDPE-B: high; Model-A: middle;

Model-B: low uniaxial strain hardening (left) and related predictions for

film width vs. DR including comparison with corresponding Newtonian

predictions by using the 2D kinematic model (right), where DR = 32.7, die

width = 250 mm and X = 90 mm. Reproduced with permission from Polym.

Eng. Sci. 41 (9), 1564 (2001) [39]. Copyright 2001, John Wiley and Sons.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589

101

Fig. 23: Prediction of steady shear viscosity (S), first normal stress difference (N1)

uniaxial (E), planar (P) and biaxial (B) extensional viscosities by using

integral constitutive equation of the K-BKZ type with PSM strain-memory

function (Eq. 56) for IUPAC-LDPE melt A at 150 °C. Solid symbols

visualize experimental data. Solid lines represent model predictions when

parameter in Eq. 56 is allowed to vary with the relaxation time, the dashed

line corresponds to the uniaxial extensional viscosity predicted by using a

single value of . Reproduced with permission from J. Nonnewton. Fluid

Mech. 97 (1), 13 (2001) [145]. Copyright 2001, Elsevier.

Th

is is

the au

thor’s

peer

revie

wed,

acce

pted m

anus

cript.

How

ever

, the o

nline

versi

on of

reco

rd w

ill be

diffe

rent

from

this v

ersio

n onc

e it h

as be

en co

pyed

ited a

nd ty

pese

t. PL

EASE

CIT

E TH

IS A

RTIC

LE A

S DO

I: 10.1

063/5

.0004

589